Sei sulla pagina 1di 99

21 Characterization of Surfaces, Interfaces,

and Thin Films of Organic Materials


David L. Allara and Ping Zhang

Department of Materials Science, Pennsylvania State University,


University Park, PA, U.S.A.

List of Symbols and Abbreviations


659
21.1
Introduction
661
21.1.1 Overview and Purpose
661
21.1.2 Analysis Requirements
662
21.1.2.1 General Requirements
662
21.1.2.2 Spatial Requirements
663
21.1.3 Analysis Probes and Probe Mechanisms
663
21.1.4 Focus of the Review
666
21.2
Survey of Characterization Techniques
667
21.2.1 Summary Tables
667
21.2.2 Photons In, Photons Out
667
21.2.2.1 Grazing Incidence X-Ray Diffraction (GIXRD)
677
21.2.2.2 X-Ray Reflectivity (XRR)
679
21.2.2.3 Spectroscopic Ellipsometry (SE)
680
21.2.2.4 Surface Plasmon Resonance (SPR)
682
21.2.2.5 Raman Spectroscopy and Surface-Enhanced Raman Spectroscopy (SERS) 683
21.2.2.6 Nonlinear Techniques: Second Harmonic Gain (SHG)
and Sum Frequency Generation (SFG)
685
21.2.2.7 X-Ray Fluorescence (XRF) and UV-Visible Fluorescence
687
21.2.2.8 Infrared Spectroscopy (IRS)
689
21.2.3 Photons In, Electrons Out
696
21.2.3.1 X-Ray Photoelectron Spectroscopy (XPS)
697
21.2.3.2 Valence Band X-Ray Photoelectron Spectroscopy (VBXPS)
and Ultraviolet Photoelectron Spectroscopy (UPS)
704
21.2.3.3 Near Edge X-Ray Absorption Fine Structure (NEXAFS)
708
21.2.4 Electrons In, Electrons Out
709
21.2.4.1 Auger Electron Spectroscopy (AES)
711
21.2.4.2 Electron Energy Loss Spectroscopy (EELS)
712
21.2.4.3 High Resolution Electron Energy Loss Spectroscopy (HREELS)
713
21.2 A A Inelastic Electron Tunneling Spectroscopy (IETS)
716
21.2.4.5 Scanning Tunneling Microscopy (STM)
and Atomic Force Microscopy (AFM)
718
21.2.4.6 Low Energy Electron Diffraction (LEED)
and Transmission Electron Diffraction (TED)
720
21.2.5 Ions or Neutrals In, Ions or Neutrals Out
722
Materials Science and Technology
Copyright WILEY-VCH Verlag GmbH & Co KGaA. All rights reserved.

658

21.2.5.1
21.2.5.2
21.2.5.3
21.2.5.4
21.2.5.5

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

Rutherford Backscattering (RBS)


Forward Recoil Elastic Scattering (FRES)
Nuclear Reaction Analysis (NRA)
Ion Scattering Spectroscopy (ISS)
Secondary Ion Mass Spectrometry (SIMS), Static SIMS (SSIMS),
Fast Atom Bombardment Spectrometry (FABS)
and Dynamic SIMS (DSIMS)
21.2.5.6 Atom Scattering
21.2.6 Neutrons In, Neutrons Out: Neutron Reflectivity
21.2.7 Dynamic Depth Profiling
21.3
Evaluation of Characterization Techniques
21.4
References

723
726
729
732

734
739
739
741
743
746

List of Symbols and Abbreviations

List of Symbols and Abbreviations


Cf
c
D(, Ex)
dp
E
Eh
Ek
F (, J
h
/
JA
k
/
M
N
n
n (co)
qz
R
T(E,E1)
x,y, z

atomic fraction of atom i


constant
efficiency factor of the detector
penetration depth
electric field
binding energy
kinetic energy
electron-optical factor
Planck constant
intensity
photoelectron current
imaginary part of refractive index
path length
mass
number
refractive index
optical function
momentum change
reflectivity
analyzer transmission function
distances, coordinates

a, (j)
P
s (co)
9
X
X (E)
fi
v
crt
cp
0
co

angles
asymmetry factor
dielectric function
angle
wavelength
mean free p a t h of a n electron
linear absorption coefficient
frequency
total photoionization cross section
angle
flux
frequency

AC
AES
AFM
DSIMS
EELS
ERD
ESCA
ESDIAD
FABS

alternating current
Auger electron spectroscopy
atomic force microscopy
dynamic secondary ion mass spectrometry
electron energy loss spectroscopy
elastic recoil detection (same as FRES)
electron spectroscopy for chemical analysis (same as XPS)
electron stimulated desorption ion angular distributions
fast atom bombardment spectroscopy

659

660

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

FRES
GIXRD
HOPG
HREELS
IETS
IRS
ISS
LB
LEED
LEIS
MIM
NEXAFS
NR
NRA
PMDA-ODA
PMMA
PS
PVP
RBS
SALI
SE
SEM
SERS
SFG
SHG
SNMS
SPR
SSIMS
STM
TED
TEM
UPS
VBXPS
XANES
XPS
XRF
XRR

forward recoil spectroscopy (same as ERD)


grazing incidence X-ray diffraction
highly oriented pyrolytic graphite
high resolution electron energy loss spectroscopy
inelastic electron tunneling spectroscopy
infrared spectroscopy
ion scattering spectroscopy
Langmuir - Blodgett
low energy electron diffraction
low energy ion scattering (same as ISS)
metal - insulator - metal
near edge X-ray absorption fine structure
neutron reflectivity
nuclear reaction analysis
poly(pyromellitiodianhydride oxydianiline)
poly(methyl methacrylate)
polystyrene
poly(2-vinylpyridine)
Rutherford backscattering
surface analysis by laser ionization
spectroscopic ellipsometry
scanning electron microscopy
surface enhanced Raman spectroscopy
sum frequency generation
second harmonic generation
secondary neutrals mass spectrometry
surface plasmon resonance
static secondary ion mass spectrometry
scanning tunneling microscopy
transmission electron diffraction
transmission electron microscopy
ultraviolet photoelectron spectroscopy
valence band X-ray photoelectron spectroscopy
X-ray absorption near-edge structure (same as NEXAFS)
X-ray photoelectron spectroscopy (same as ESCA)
X-ray fluorescence
X-ray reflectivity

21.1 Introduction

21.1 Introduction
21.1.1 Overview and Purpose

Organic materials, once considered to be


of minor importance compared to metals,
ceramics and other traditional inorganic
materials, are now key components of
many contemporary materials structures
appearing in applications ranging from
microelectronics to biomedicine with the
role of the organic component ranging
from protective coatings to load bearing
components. The premier organic materials are polymers but, in addition,
monomeric molecular materials such as
liquid crystals and even surfactants are
finding many technological applications as
components in complex structures and
these structures have become the subject of
a number of scientific studies. In a large
number of applications the surfaces and
interfaces of component materials play a
dominant role in determining the useful
properties of the material item. This situation certainly pertains to organic materials
and presses the point that high quality
characterization of surfaces and interfaces
is necessary for proper utilization and development of these materials. In principle,
this would appear not to present a problem
since for years an enormous amount of
effort has gone into the development of
very sophisticated techniques for characterization of the surfaces and interfaces of
common inorganic materials such as
metals and semiconductors and it would
seem that these techniques could be carried
over to organics. However, it has turned
out that many of the useful techniques developed early on for the inorganic materials have not been either directly applicable
or very informative for organic materials.
To deal with this problem, some of the
traditional techniques have been modified

661

and some new techniques have been developed. In fact, because of striking advances
in instrumentation and electronics, a proliferation of new techniques and variations
of old techniques has arisen. In addition,
major facilities such as synchotrons and
neutron reactors have become more common and accessible for surface analysis.
New types of lasers continually add possibilities to the growing list of surface techniques and powerful computational tools
now make complex data analysis standard.
The choices of techniques to apply to surface/interface/thin film characterization of
organic materials seems somewhat bewildering, especially in view of extreme ranges
of cost differences for obtaining similar information by different techniques. The objective of this review is to provide the reader with both a general understanding of the
requirements often arising in analyzing organic surfaces and the types of tools which
have proven useful to this task. Hopefully
this information will serve as a guide as to
which techniques are optimal for a given
analysis requirement.
This chapter is organized as follows.
Section 21-1 describes the general requirements for surface and interface analysis of
organic materials, classifies the types of experiments on the basis of particle in-particle out categories and briefly discusses the
analysis interaction mechanisms of the
probes. Section two presents detailed discussions of each analysis method, presented in an order based on the particle i n particle out classification. The material is
summarized in several tables at the beginning of the section. Finally, in section three
an attempt is made to evaluate the optimal
usefulness of the techniques presented in
Section two on the basis of a number of
criteria common to many laboratories.

662

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

21.1.2 Analysis Requirements


21.1.2.1 General Requirements

In general, the major difference between


inorganic and organic materials is that in
the latter, the constituent atoms are bonded together in units recognized as "functional" groups and the overall properties
of the material derive from the interactions
and arrangements of these functional
groups in a manner conceptually parallel
to the atomic interactions in typical inorganic materials. Likewise, the surfaces and
interfaces of organic materials are viewed
in terms of the arrangements of these
molecular groups. Since the latter often
consist of several atoms, e.g., CO 2 H, even
the orientation of the group at the surface
can effect more subtle properties such as
chemical reactivity and biological response. There are three aspects of these
surfaces which are highly relevant to designing analysis strategies: surface stability, molecular complexity and environmental effects. Corresponding to these aspects
it is important to identify analysis techniques which would be nondestructive,
provide high information content and function in situ. Each of these latter three aspects is discussed briefly below.
Organic materials have been sometimes
informally referred to by physicists as
"soft" condensed matter. This term aptly
suggests that these materials often exhibit
low melting or softening temperatures as
well as low hardness, for example. Of
course, these types of properties are the
outcome of the much weaker forces usually encountered in the interactions between
molecules or macromolecules in organic
materials as compared to the stronger
binding forces encountered in the ionic
and covalent interactions of most inorganic materials. One of the manifestations of
this situation is that the surfaces of organic

materials are often very dynamic in the


sense that both the detailed, but important, arrangement of constituent molecular groups exposed at the surface and the
surface morphology often can vary with
introductions of modest energies, for example, from small upward excursions in
temperature, from solvation energies arising from the presence of a contacting liquid or from the interaction with particle or
photon beams. This characteristic is obviously of serious concern in devising surface
and interface analysis strategies since the
surface analyzed should ideally not be
modified by the analysis technique itself.
One marvelous attribute of organic materials is the seemingly endless variation of
combinations in which organic molecules
can be formed from just a few common
elements, in particular C, H, O and N.
While this provides opportunities to tailor
material properties via small changes in
composition, conformational sequences or
isomeric structures, for example, it also introduces strong demands on the information content of surface analysis techniques.
For example, a technique may be called
upon to provide quantitative details of
subtle molecular structure features responsible for surface properties such as wetting,
bio-activity or friction.
A final aspect of organic surfaces is the
occasional need to carry out analyses in
the presence of liquid solutions. Such in
situ analyses provide extreme restrictions
on techniques and eliminate all vacuumonly methods such as those using beams of
charged particles, e.g., Auger spectroscopy
and secondary ion mass spectrometry.
However, the need for in situ analyses is
extremely important in applications involving bio-active surfaces such as those in
biomedical implants exposed to biological
fluids. Another example is the analysis of
the growing polymer surface during the

21.1 Introduction

electropolymerization of monomers, such


as aniline, to form conducting polymer
film electrodes.
21.1.2.2 Spatial Requirements
What makes a surface analysis different
from a bulk analysis is simply that in the
former, information is obtained from a selected region of the sample. In a more general sense then one should approach the
subject of surface and interface analysis
from the point of view of spatial selectivity.
Techniques that exclusively derive information from the surface region are termed
surface selective. It is obvious that in order
to properly classify the technique the actual size of the region should be specified.
This is best done by specifying the depth
away from the surface past which no information effectively is obtained. In the case
of analysis of an interface one is primarily
concerned with selecting a region between
two phases. Because the absolute amount
of material to be analyzed can become
vanishingly small as the analysis region becomes more and more localized, the signalto-noise ratio capabilities of the measurement technique become of critical concern.
The term surface sensitivity is appropriate
to this aspect. An analysis technique which
obtains information from a region including, but much deeper than, the surface can
be surface sensitive but not surface selective
provided only that the surface information
portion be above the signal-to-noise ratio.
Most techniques that depend exclusively
upon photon probes fall into this category.
Most techniques that depend upon analyzing emitted charged particles from the surface are surface selective. Very few techniques are truly interface selective. One example is sum frequency generation, a second order, non-linear optical response
technique which depends upon the breaking of structural symmetry at an interface.

663

An important refinement of surface selectivity is the ability of a technique to


measure information gradients away from
a surface to the end of the analysis region.
This is usually referred to as depth profiling
and can both be done nondestructively in
cases for which the analysis mechanism allows extraction of this information and destructively by simply sequentially removing layers of material (usually by sputtering with ion beams) and analyzing the
newly exposed surface.
Another important spatial aspect is lateral resolution. This term refers to obtaining information from selected, localized
spots across a surface or interface. The size
of the spot can vary from atomic dimensions, as in scanning tunnelling microscopy, to the mm scale, typical of many methods. If the information can be connected to
give a continuous picture of the surface,
the technique is then an imaging one.
21.1.3 Analysis Probes and Probe
Mechanisms
It is convenient to classify probes as either macroscopic or microscopic. The former refers to measurements which depend
upon a macroscopic-scale response of the
surface such as wetting, friction or chemical reactivity. Of course, these properties
are ultimately based on the microscopic or
molecular nature of the material but very
complex models of the mechanisms are required to extract this microscopic information from the macroscopic measurement,
e.g., the shape of a drop of liquid placed on
the surface. On the other hand, the interaction of the surface with a beam of light can
produce a spectroscopic measurement
which only can be understood by first
using a model based on the quantum
mechanical responses of the atoms or
molecules in the surface region. Probes
which operate primarily by highly local-

664

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

ized interaction mechanisms are thus microscopic. Such probes consist primarily of
photons, ions, atoms and polyatomic species, electrons and neutrons. It is these
atomic scale probes which will be the subject of this review. Since the details of the
characterization experiment, the capabilities of the method and the interpretation of
the data each depend directly on the mechanisms of interaction with the sample, it is
convenient to review some of the more pertinent details of classification here. Further
details for each probe can can be found
under the corresponding analysis technique in Sec. 21.2, both explicitly and in a
number of the references given.
Probe mechanisms can be classified into
three overall types: (1) the disappearance
of an input particle (absorption) or appearance of an output particle (emission)
as the sole event, (2) the change of state of
an input particle upon collision (scattering,
both elastic and inelastic) and (3) the ejection of a new particle as a result of a different input particle (sputtering, reactive scattering, desorption or stimulated emission).
Each of these types is discussed below in
order and schematic representations are
given in Fig. 21-1.
Because the probes and the confining
volumes of the interaction regions are of
atomic dimensions, the appropriate models for the interaction mechanisms are
quantum mechanical. However, in some
cases classical models work quite well, e.g.,
Rutherford back scattering. The type of
mechanism which becomes operative upon
collision of a probe particle with a target
atomic or molecular solid depends primarily upon the match between the energy of
the input probe and the associated energy
spacings of the states of the target particles
which will be involved in the probe-target
interaction. Thus it is important to understand that the entire mechanism of a par-

(a)

ABSORPTION

EMISSION

(b)

SCATTERING

(c)
COLLISIONEJECTION

Figure 21-1. Schematic representation of different


types of probe mechanisms, (a) The disappearance of
an input particle or appearance of an output particle;
(b) the change of state of an input particle upon collision and (c) the ejection of a new particle.

ticular type of experiment, e.g., ion scattering (ions in-ions out), may change just by
changing the energy state of the incoming
probe.
The simplest experiments, at least conceptually, are absorption and emission.
With the exception of transmission electron microscopy and high voltage-field
emission tip imaging, the only experiments
of concern are those which involve onephoton processes. The latter are carried
out by either: 1) sending in light and measuring the fractional absorption or 2) measuring light emission induced by a macro-

21.1 Introduction

scopic form of excitation such as thermal


energy (temperature). If the information
gathered is sorted out on the basis of response per energy or wavelength interval
of the light then absorption or emission
spectra are obtained. The useful wavelength region for surface and thin film absorption spectroscopy ranges from the uv
to the far-infrared. Emission is rarely done
and then only at infrared frequencies. For
simplicity, emission of light is classified as
a photon in-photon out process.
In a scattering experiment, both momentum and energy can change upon collision of the probe particle and this change
can be related to structural characteristics,
often in a quantitative manner. In elastic
scattering there is a change of momentum
but no energy change. The best example of
this is diffraction, an experiment in which
the angular distribution of the elastically
scattered particles is related to intrinsic
spacings of the scattering centers in the
surface region of interest. Generally, precise diffraction results require the wavelength of incoming particle to be of the
same magnitude as the latter spacings.
These conditions effectively can be met for
the spacings in solids by the use of X-ray
photons, typical reactor-produced neutrons, high voltage-accelerated electron
beams or thermal velocity He-atom beams,
all of which exhibit Angstrom-scale wavelengths. Since the success of a diffraction
experiment depends upon the signal to
noise ratio for the diffracted beam intensities, each of which carry only a fraction of
the scattered particle intensity, it is important to provide a large input particle flux,
particularly for surface layer scattering
which involves at a minimum only one
monolayer of scattering centers. Accordingly, diffractive scattering from surfaces
usually involves high flux sources such as
synchotrons, very sensitive, low-noise de-

665

tectors and long counting times. A simpler


experiment is reflectometry, which involves precise measurements of the angular dependence of the specular reflectivity
from a beam of collimated, monoenergetic
particles, almost always neutrons or Xrays, reflected from a thin film structure in
order to determine physical aspects of the
film and interface structures. Ellipsometry,
a class of closely related experiments, involves the specular reflection of monochromatic, optical frequency light from a surface in which measurements of the relative
changes of the phase shifts and amplitudes
of two polarization states carried by the
input beam can yield information about
structure in thin film samples. All of these
reflection experiments essentially depend
upon the momentum properties of the
light as manifested in interference and refraction effects. However, in cases of optical frequency light, absorption will occur if
appropriate electronic states are present in
the material. In general, when scattered
probe particles, of any type, exhibit
changes in energy (frequency for photons
or kinetic energy for matter) spectroscopic
information results which can be interpreted in terms of structure in the surface region of the sample with this interpretation
dependant upon the types of energy exchange processes occurring in the scattering volume. These latter processes will
consist of quantum excitations in the solid
although sometimes simple classical momentum exchange pictures are accurate
approximations to the mechanism.
Collision-ejection processes are usually
the most complicated mechanisms of surface probes because of the involvement of
two or more different types of particles.
Further, in experiments in which atomic or
molecular fragments of the target material
are ejected, the incoming particle is required to deposit considerable kinetic en-

666

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

ergy to sputter off the fragments and accordingly a disruption of local structure
throughout the collision volume occurs.
The general types of input-output pairs of
particles useful in surface and interface
analysis involve photons-electrons (photoelectron emission), electrons-ions and
ions (or neutrals)-ions (or neutrals). Nonlinear optical processes should be considered as collision-ejection processes rather
than scattering because the interaction
mechanisms are explained rigorously on
the basis of the annihilation of the input
photon(s) and the creation of the output
photon(s), each of which usually have different energies and momenta and are considered as different (discrete) photons.
Mechanisms involving photons require
quantum mechanical models, and often
electrodynamical as well, in order to explain properly photon-matter interactions. For example, photoemission experiments involve excitation of electrons from
bound states in the solid into vacuum continuum states, a quantum process. However, nonphoton collision-ejection processes
often can be understood reasonably well in
terms of classical or semiclassical collision
mechanics.
21.1.4 Focus of the Review
The above paragraphs have defined
briefly the essential factors to be considered. A broad coverage of the literature
has been made to find examples of analytical techniques applied to characterization
of the surfaces and interfaces of organic
materials. The examples consist primarily
of polymers since these are by far the most
common organic materials, but some examples covering thin molecular (surfactant
type molecules) films are also given. The
techniques have been limited to those using atomic-scale probes and have been cat-

egorized in terms of the probe particle


bombarding the sample and the corresponding ejected or scattered particle
which is analyzed in some sort of detection
system for relevant information about the
desired sample characteristics. Thus infrared absorption spectroscopy consists of an
IR photon in and an IR photon out whereas X-ray photoelectron spectroscopy consists of an X-ray photon in and an electron
out. After the particle in-particle out classification the techniques are further subdivided into those that: (1) probe a region
bounded by the outer surface (towards the
incoming probe) and existing to a depth of
5-10 nm or less from this surface and (2)
probe a region greater than 5-10 nm but
generally no deeper than about 1 jim from
the surface, although a few techniques will
probe bulk samples, depending on the experimental mode, e.g., transmission infrared spectroscopy. With respect to the latter
category, the selection criteria for inclusion of a specific technique is to consider
bulk analysis techniques, viz, those having
no surface selectivity, only if they demonstrate surface sensitivity at the near-monolayer level or better. Each technique is discussed in sufficient detail to indicate its
ability to meet the analysis requirements
given in Sec. 21.1.2 above. It is, of course,
implicit in all analyses, regardless of which
particular sample characteristic is being
measured, that the measurement be as
quantitative as required and thus the
aspect of quantitation runs constantly
through the presentation of the techniques. Tables are presented as a convenient summary of the more detailed information given in the text. As is common in
the field of surface characterization, the
various techniques are given acronyms for
convenience. The ones adopted for this review are summarized in the List of Symbols and Abbreviations. They have been

21.2 Survey of Characterization Techniques

selected to conform with the most common usage wherever this can be clearly determined. A number of texts emphasizing
different aspects of surface analysis and
associated instrumentation have been written and are listed below. The reader is referred to them for general background
reading.

21.2 Survey of Characterization


Techniques
21.2.1 Summary Tables
For convenience in viewing and comparing the capabilities of the large number
of analysis techniques, the survey has been
summarized into several tables. One of the
most important concerns in a surface type
of analysis is that of understanding exactly
which depth regions of the surface will be
probed, i.e., what is the surface selectivity.
A natural break occurs for most of the
techniques at the 5-10 nm analysis depth
and Tables 21-1 and 21-2 have divided the
techniques on this basis. Table 21-1 covers
techniques that are selective for the top
5-10 nm or less of the surface, i.e., these
techniques exhibit the sensitivity to give
reliable information for films or surface
regions in this range. However, because of
this surface selectivity it should be realized
that these techniques also are limited to
this thickness range and cannot give information from deeper levels without removing (sputtering, etc.) surface layers. On the
other hand Table 21-2 covers techniques
which obtain information from greater
depths simultaneously with the top surface, and thus surface region information
is contained in the total signal. In some
cases the surface information can be extracted from the latter by taking advantage
of theoretical relationships between signal

667

characteristics and the depth region of


probe-sample interactions, provided, of
course that the surface-region sensitivity is
sufficient to ensure that the contributing
signal is above the noise level. Thus while
the latter techniques are not surface selective they are usually surface sensitive. Further, although Table 21-2 is titled as describing techniques probing < 1 |im, some
of them will analyze even deeper regions,
in the limit becoming bulk analyses. Finally, leading references are included with the
tables so that the reader quickly can get
detailed information on the subject without referring to the related section in this
review. To give the reader a concise comparison of relative capabilities of the techniques, Table 21-3 presents a summary of
selected advantages and disadvantages of
the techniques. One of the entries of this
table is the usable environment which is
possible for analysis, an important factor
in analysis for applications such as those
involving biological or electrochemical environments.
21.2.2 Photons In, Photons Out
Photons have been the basis of traditional spectroscopic and diffraction methods for the solid state. As a consequence,
there is a long history of well-understood
relationships between the structure of matter and its behavior with light beams. Accordingly, it is a great advantage to adapt
these photon interactions to the problem
of surface and interface analysis wherever
possible. However, relative to ions and
electrons, the most common surface
probes, photons provide some complexities in devising experimental strategies for
surface analysis. This situation can be understood in terms of the fact that electron
and ion beams can be generated at a wide
variety of energies using essentially the

Molecular structure
orientation

Atomic composition

Molecular structure

Ordering

Topography;
electronic states

SFG

XPS

VBXPS

NEXAFS

UPS

AES

HREELS

LEED

STM/AFM

ISS

SSIMS

He
scattering

(visible+ IR)/
(visible+ IR)

X-ray/e"

X-ray/e~

X-ray /e~

UV/e~

e~/e~

e-/e-

e-/e-

e"/e-

ion/ion

ion/ion

atom/
atom

Ordering

Molecular structure;
composition

Atomic composition;
molecular structure

Molecular structure

Molecular structure;
orientation

Molecular structure

Atomic composition;
chemical group
identification

Molecular structure
orientation

SHG

visible/
visible

Information
obtained

Name

Probe
in/out

~mm

0.3-0.5 nm

~ 1 nm

0.3-0.5 nm

Semiquantitative
No

mm

Semiquantitative
> 50 nm

> 150 urn

Yes

Not possible
currently

No

10" 5 -10~ 6

No

Yes

10" 2 -10~ 3

Monolayer

No

Single
atom

No

No

Submonolayer

>0.1 nm

Monolayer

Yes, with variable


take-off angle

Semiquantitative

10-2-10"3

> 5 nm

Yes, with variable


take-off angle

Submonolayer

Difficult

Yes, with variable


take-off angle

Yes, with variable


take-off angle

No

No

Depth profile
capability

~mm

Submonolayer

io- 2 -io~ 3

Submonolayer

Submonolayer

Sensitivity
at.%

Submonolayer

Difficult

Yes

Semiquantitative

Yes

Quantitative

~mm

100 um-mm

100 um-mm

>10 um

>10|xm

Lateral
resolution

Atomic scale
Atomic
surface layer

~ 1 nm

~ 1 nm

<10 nm

~ 0.5 nm

~10nm

0.5-10 nm

0.5-10 nm

Interface
sensitive

Interface
sensitive

Depth
resolution

Camillone III
etal. (1993 a)

van der Wei


et al. (1990)

Gardella and
Pireaux (1990)

Frommer (1992)

Rous et al. (1990)

Pireaux et al. (1990 a)


Dubois (1993)

Ramaker (1991)

Salanek (1985)

Jordan - Sweet
(1990)

Salanek (1985)

Powell and Seah


(1990)

Bain et al. (1991)


Vogel and Shen (1991)

Zhang et al.
(1992)

Selected review
or papers

EJ

0
0)

CO

e rfaces, and 7hin Fil

(Q
Q5

) SOI

Table 21-1. Techniques that probe only the top 5-10 nm.

laracterization 0"
mic Materials

21.2 Survey of Characterization Techniques

669

Table 21-2. Techniques with surface selectivity up to 1 um.


Probe
in/out

Name

Information
obtained

Depth
resolution

X-ray/
X-ray

GIXRD

Molecular
orientation;
crystal structure

nm

cm

No

X-ray/
X-ray

XRR

Concentration
profile;
interface width

lnm

cm

Yes

X-ray/
X-ray

XRF

Concentration
profile

1 nm

cm

UVvisible/
UVvisible

SE

Film thickness;
concentration
profile; interface
width

-lnm

visible/
visible

Raman

Molecular
structure

visible/
visible

SPR

Film thickness;
morphology

IR/IR

IRS

Molecular
structure and
orientation

X-ray/
e

XPS& Concentration
sputter- profile; some
chemical
ing
information

Lateral Quantiresolutative
tion

Depth probed

Selected review or papers

Surface selective
< 10 nm for critical angle mode

Factor et al.
(1991)

Monolayer

- 300 nm

Russell (1990)

Yes

Monolayer

Surface selective
< 1 0 nm for critical angle mode

Bohn and
Miller (1991)

mm

Yes

Submonolayer

Surface selective
< 1 0 0 n m for
critical angle
mode

Collins et al.
(1993)

~l-20nm

1 um

Not
usually

>
Monolayer

1 nm

- 1 0 um

Yes

Submonolayer

-50nm1 um

- 1 0 jim

Yes

Submonolayer

>10nm

100 u m 10 mm

Yes

Yes

EELS

Atomic
composition

100 nm
Thin film

5 um

e7e"

TED

Ordering

<10nm
Thin film

Near
atomic

ion/on

NRA Atom concentration profile;


interface width

-20nm

mm

Yes

ion/
7-ray

NRA

-lOnm

mm

Yes

ion/ion

FRES Atom concentration profile;


interface width

-1080 nm

mm

Yes

-30nm

mm

Yes

ion/ion

RBS

Elemental concentration profile;


elemental
identification

io- 2 -

Surface selective
Bohn and
< 1 0 n m for surface Walls (1991)
enhanced mode
- 200 nm

io- 2

DeBruijn
et al. (1991)

Surface selective Allara (1993)


< 1 um for critical
angle mode
um

Schwamm
et al. (1991)

<100nm

Briber and
Khoury (1988)

10~3

e-/e"

Atom concentration profile;


interface width

Sensitivity
at.%

Strong and
Whitesides(1988);
Dorset (1991)

Monolayer

io- 2 io- 3

(am

Chaturvedi
et al. (1990)

io- 2 io- 3

- um

Endish et al.
(1992)

io- 2 io- 3

um

Sokolov et al.
(1989 a)

io- 2 -

um

Green and
Dogle (1990)

10~ 3

670

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

Table 21-2. Continued.


Probe
in/out

Name

Information
obtained

ion/ion DSIMS Concentration


profile;
chemical
and elemental
neutron/
neutron

NR

Concentration
profile;
interface width

Depth
resolution
~13nm

4 nm

Lateral Quantiresolu- tative


tion

Sensitivity
at.%

Depth probed

Selected review or papers

-'jim

Semiquantitative

10 5 10'6

~ urn

Russell et al.
(1989);
Zhao et al.
(1991a)

' cm

Yes

> Monolayer

~ 300 nm

Russell (1990)

Table 21-3. Major advantages and disadvantages of surface and interface analysis techniques for organic materials.
Selected advantages

Selected disadvantages

Usable
environment

Molecular, atomic ordering

Strict sample requirements


Sample damage
Synchotron usually required

Air and vacuum

XRR

Nondestructive depth profile


High precision film thickness
and morphology
Excellent depth resolution

Data interpretation difficult,


simulation needed
Strict sample requirements
Other experiments needed
to complement results

Air and vacuum

SE

Nondestructive depth profile


High precision film thickness
and morphology
Electronic structure
Time resolved

Data interpretation difficult,


simulation needed
Other experiments needed
to complement results

Liquids, air
and vacuum

SPR

Nondestructive
High precision film thickness
and morphology
Time resolved

Data interpretation difficult,


simulation needed
Strict sample requirements

Liquids, air
and vacuum

Raman

Molecular structure
Chemical bonding

Poor surface sensitivity


Usually not quantitative

Liquids, air
and vacuum

SERS

Molecular structure
Chemical bonding
Interface selective

Highly specialized samples required

Liquids, air
and vacuum

SHG

Molecular orientation
Nondestructive
Interface selective
Time resolved

Interpretation requires considerable


knowledge of sample

Liquids, air
and vacuum

Name

GIXRD

671

21.2 Survey of Characterization Techniques

Table 21-3. Continued.


Name

Selected advantages

Selected disadvantages

Usable
environment

SFG

Chemical bonding information


Nondestructive
Interface selective
Can be quantitative
Time resolved

Difficult experiment

Liquids, air
and vacuum

XRF

Surface selective
Nondestructive
Accurate depth profiling

Strict sample requirements

Liquids, air
and vacuum

Surface selective
Nondestructive
Depth profiling

Strict sample requirements

Liquids, air
and vacuum

Excellent chemical information


Nondestructive
Can be quantitative

Poor depth resolution

Liquids, air
and vacuum

UV-visible
Fluorescence
IRS

XPS

Surface selective
Chemical information limited by
Chemical group information
instrument resolution
Easy quantification
Sample charging for monochromatized
Little sample damage
X-ray source
Easy sample preparation
Depth profiling with variable angle

Vacuum

VBXPS

Surface selective
Chemical bonding information
Structure information
Little sample damage
Easy sample preparation

Spectral interpretation difficult, MO


Calculations needed
Long data acquisition time
Quantification difficult

Vacuum

UPS

Very surface selective


Chemical bonding information
No sample damage

Spectral interpretation difficult, MO


calculations needed
Quantification difficult

Vacuum

Surface selective
Chemical bonding information
Molecular orientation information

Tuneable X-ray source necessary


(synchotron)

Vacuum

AES

Surface selective
Imaging capability
Excellent lateral resolution

Severe sample damage


Severe surface charging

Vacuum

EELS

Surface selective
Quantitative
Sensitive to low mass elements

Sample damage
Sample preparation difficult

Vacuum

HREELS

Surface selective
Chemical information
Highly surface sensitive

Quantification difficult
Severe surface charging

Vacuum

IETS

Chemical information
Highly surface sensitive

Cyrogenic condition needed


Sample preparation difficult
Application very limited

Liquid He

NEXAFS

672

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

Table 21-3. Continued.


Name

STM/AFM

LEED
and TED

Selected advantages

Selected disadvantages

Extremely surface selective


Atomic resolution imaging of surface
topography

Imaging mechanisms not well


understood - quantitative interpretations can be difficult
STM can't image insulators
Sample damage to polymers
and organic films possible
Chemical sensitivity poor

Molecular and atomic ordering

Sample damage
Conducting substrate required

Usable
environment
Liquid, air
and vacuum

Vacuum

RBS

No chemical information
Generally nondestructive depth profile
Easy quantification
Some sample damage is possible
Absolute concentration determined Low sensitivity for low mass elements
Data interpretation straightforward

Vacuum

FRES

Excellent for H and D detection


Generally nondestructive depth profile
Easy quantification
Absolute concentration determined
Data interpretation straightforward

Only limited elements detected


Relatively poor depth resolution
Some sample damage

Vacuum

NRA

Nondestructive depth profile


Very high depth resolution
Absolute concentration determination

Only limited elements detected


No chemical information
Long data acquisition time

Air

ISS

Highly surface selective


Structure information possible

No chemical information
Quantification difficult
Some sample damage

Vacuum

SSIMS

Imaging capability
Highly surface selective
Chemical structure information
High sensitivity

Quantification difficult
Spectrum very complicated
Surface charging

Vacuum

DSIMS

Highly surface selective


High sensitivity
Depth profile capability

Sample damage inherent


Relatively poor depth resolution because
of artifacts

Vacuum

Atom
scattering

Highly surface selective


Molecular ordering

Difficult experiment
Strict sample requirements
Neutron source facility required
Long acquisition times
Deuterated materials often required

NR

Nondestructive depth profile


Excellent depth resolution

same principles, i.e., acceleration of


charged particles by electric (or magnetic)
fields, whereas the generation of light,
from X-rays to microwaves, can utilize a
variety of sources including synchotrons,
lasers, hot filaments and klystrons. In

Air

many experiments involving both photons


and matter, the output beams exhibit a
range of energies (frequencies) and must be
resolved into groups of particles within
small, given energy ranges (resolution elements) in order to perform the analysis.

21.2 Survey of Characterization Techniques

For charged particles this is done with manipulation by electric fields but for photons several methods are used ranging
from diffraction gratings to Michelson interferometers, the latter being the basis of
Fourier transform spectroscopy, and each
method has a number of unique variations
depending upon the frequency range. Finally, detection of a charge is a much easier
feat electronically than detection of a photon and, in fact, for very high sensitivity,
photons are first converted to electrons
which are then detected. Based on these
contrasting factors, purely electromagnetic
probes have found a valuable place in surface analysis when they can be adapted
appropriately. In order to understand
these adaptations to surfaces it is necessary
to understand the basic nature of the general experiment.
For analytical purposes, nearly all experiments with light can be understood at
a useful level on the basis of two types of
theory: (1) the quantum theory of radiative
excitation and decay between energy states
and (2) the classical theory of the interaction of electromagnetic waves with dielectric media. The first involves spectroscopic
resonances between the frequency of the
light and the energy separations between
quantum levels in the solid and the intensities of the resonances are characterized in
terms of cross sections or probabilities of
the associated quantum excitations. Spectroscopic information reveals associated
structural features which set the positions
and numbers of the energy states, e.g., the
types and symmetries of chemical bonds as
revealed by vibrational spectroscopy or
the number of conjugated Ti-bonds as revealed by electronic absorption. The second takes account of the wavelength (momentum) character of the light and treats
the phase interference effects which occur
upon reflection at the interfaces between

673

different media as light is propagating.


Diffraction experiments are explained on
the basis of a discontinuous medium of
scattering centers periodically spaced at
distances comparable to the wavelength of
the light; for example, atoms in a crystal
lattice scatter ~ angstrom scale X-rays.
However, for longer wavelength light,
such as in the visible region, uniform density materials appear continuous with average properties to these probing wavelengths and the optical response can be
understood in terms of material parameters such as refractive indices and electromagnetic susceptibilities. At typical (low)
levels of radiation power the optical response of a material, e.g., reflectivity from
a smooth surface, depends upon a linear
response to the electric field amplitude and
can be described quantitatively by means
of a frequency dependent refractive index
value, or equivalently a dielectric function,
assigned to the medium of interest. In a
general case, the refractive index and the
equivalent dielectric function are defined
as frequency-dependent complex quantities:
n (co) = n + i k

(21-1)

s(co) =

(21-2)

s1-\-is2

These two quantities are related by the


complex equation
n(co)Y = e(co)
(21-3)
With respect to #(co), the real part accounts for effects such as refraction and
the imaginary part for absorption (excitation or loss). The physical basis of the
dielectric function is similar. The inclusion
of an absorption term allows one to deal
with spectroscopic experiments in which
the medium (media) contains sets of resonant energy states and thereby absorbs
light at the corresponding frequencies. At

674

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

high intensities of light (with lasers) the


optical response will change to include behaviors dependent upon the square and
higher powers of the electric field amplitude which can include the simultaneous
combined intensities of different photons
of different frequencies. These nonlinear
responses give rise to a variety of very elegant spectroscopies, some of which are
useful for surface analysis, as will be discussed below.
The surface selectivity of a photon
probe is determined both by the ability to
confine the overall distribution of the electromagnetic fields of the interacting light
in the surface regions of the target solid
and by the ability to further localize the
interaction mechanisms of these existing
fields to an even more specific surface region. The first mechanism is fairly universal while the second is highly specialized
and depends upon nonlinear responses of
the material. In general, for a medium with
a smooth, planar surface a useful rule of
thumb is that the electromagnetic fields
cannot be localized at the surface much
better than within a depth equal to a major
fraction of the wavelength of the probing
light. In the simplest experiment, propagation of light through a medium, the light
will continue on infinitely into the medium
unless it is absorbed, so there is no surface
selectivity. If light is reflected from a surface, a portion of the wave is also transmitted into the medium and since the reflection and transmission are coupled, i.e.,
they are one event together, the behavior
of the transmitted beam deep in the medium, e.g., another reflection or absorption
by some species, will affect the reflection.
This can be described in terms of Fig. 212 a where a beam of intensity Io impinges
on the boundary between materials 1 and
2 to generate a reflected beam of intensity
/ 1R at the exit angle (j> and also a transmit-

(a)
1T

(b)

<E 2 :

(c)

Figure 21-2. Schematic representation of the surface


selectivity of a photon probe.

ted beam with intensity / 1 T . The diagram


shown implies that the refractive index ratio n2/n1 > 1 since the transmitted beam in
medium 2 refracts to a steeper angle to the
surface normal than the incoming angle (j)
(Snell's law). Thus there is no special surface selectivity in simple reflectivity and as
a consequence a simple reflection spectroscopic measurement on a bulk solid, e.g.,
infrared reflection spectroscopy of a sheet
of polymer, only can be useful for surface

21.2 Survey of Characterization Techniques

analysis by first subtracting a large bulk


signal. However, there are classes of experiments in which the light can be guided
along the boundary between two adjacent,
planar media with the result that the
bounded electromagnetic fields, termed
evanescent waves, can be constrained to
penetrate into the lower refractive index
medium to a depth slightly less than the
wavelength of the light. This is shown
schematically in Fig. 21-2 b where the light
reflects off the boundary between media 1
and 2. In this case n2/n1 > 1 as in Fig. 212 a but the beam is propagating through
medium 2 now. When the angle of incidence (j) is greater than a critical value </>c,
the transmitted intensity 71T = 0 and the
electric field of the light decays exponentially away from the phase boundary as
shown schematically in the figure by the
shading. In Fig. 21-2 c a plot of the mean
square value of the electric field against
distance is shown. A general equation for
this distance dependence is
<E2Z)

(21-4)

where < 2 >/< 2 0 > is the ratio of the mean


square value of the electric field at a distance z away from the phase boundary interface to the value at the interface, c is a
collection of constants and dp is the penetration depth defined as
.

I Hi

sin m

(21-5)

where X is the wavelength of the incoming


light, ni and n2 are the refractive indices
(usually always real values) of the adjoining media and 4> is the angle of incidence.
For such configurations one can arrange
to have the lower index medium be the one
of interest and thus some depth region into
the medium can be probed selectively. Ex-

675

periments which depend upon evanescent


waves include XRF, SPR, waveguide Raman and internal reflectance IRS. These
range from the X-ray to the infrared region
and the surface selectivity has been observed correspondingly over a scale of
~10nm to ~ 1 [im. However, it is very
important to realize that these experiments
are highly material dependent and thus
many imagined specific experiments cannot be done because the sample of interest
does not have the necessary optical properties for the wavelength region desired. Out
of the many combinations possible only a
relatively small number of useful examples
have been reported and in particular, the
X-ray experiments are very limited although these obviously would be the ones
most desired because of the nm-scale surface selectivity possible.
Although, as the above discussion
points out, there is no specific surface selectivity because of the tendency of light to
be distributed throughout a sample, it is
possible to take advantage of the extreme
sensitivity of this distribution to the structure of the sample and make quantitative
determinations of the structural features
by measurements of the reflectivity and/or
transmissivity of the light beam with the
sample. The principle is simple and essentially depends upon two fundamental
properties of waves propagating through
parallel layer media. The first property is
the inherent tendency of the beam to partition some of its intensity between reflection and transmission at every interface or
material phase boundary in the sample, as
was schematically illustrated in Fig. 21-2 a.
In any optical experiment, the exact degree
of partitioning, / R // o and / T // o depends
upon the refractive index difference between the two adjoining phases, the wavelength of the light, the angle of incidence
and the polarization direction of the input

676

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

beam. The second property is that of interference. This phenomenon is simply described with reference to Fig. 21-3, which
depicts a beam of light entering a thin film
(phase 2) supported on an infinite substrate. The complete interaction of the
beam with the sample is understood in
terms of a standing wave permeating the
entire structure. Within the film layer, it is
possible to picture the interaction in terms
of the beam bouncing within the phase
boundaries to create constructive and destructive interference effects which regulate the intensity of the total output light
beam. The exact outcome of these interference effects depends upon the refractive
indices of the media and the thicknesses of
the film. Thus if the experimental variables
are fixed, viz, angle of incidence and wavelength, the total reflectivity can be related
directly to the layer thickness and refractive indices. A similar picture holds for
multiple layers. These types of measurements and relationships are the basis of
XRR, SE and several types of IRS experiments and the general principles underlie
the quantitative interpretation of any experiment in which light interacts with condensed media. It should be noted that the
basic equations used to interpret data ap-

Figure 21-3. Schematic representation of the reflection of light from a multiple-layer medium.

ply to all wavelengths, from X-rays to radio frequencies.


The surface sensitivity of an experiment,
by the definition in this review, depends
completely upon the signal-to-noise ratio
of the measurement with regard to the
fraction of the total signal contributed by
the surface region of interest. For almost
all of the photon in-photon out techniques discussed below, the signal-to-noise
ratio is sufficient in favorable cases, with
the best instrumentation available, to detect the amount of signal due to a surface
layer, viz, a monolayer. In many cases
fractions of a monolayer can be easily detected. As expected, the problem often is
simply that with no corresponding surface
selectivity the surface signal contrast is
low, i.e., the bulk signal obscures the surface contribution.
One general advantage of the photon
techniques is the fact that photon beams
tend to be much less destructive towards
sensitive organic surfaces than ion and
electron beams. This rule generally can be
understood in terms of the fact that photons possess much different momentum
properties than matter and can't remove
material by billiard-ball type collisions as
occurs in sputtering by atoms or ions.
Thus unless the photon is absorbed by
some quantum excitation process it will
pass through the material with no energy
dissipation. However, even in the case of
materials which absorb light, most of the
energy is dissipated rapidly away from the
local area in the form of heat and typical
beam intensities do not heat the material
sufficiently to cause damage. Exceptions
to this rule occur with extremely high flux
synchotron-generated X-rays for surface
scattering, high powered laser beams for
nonlinear spectroscopies and materials
which are photochemically sensitive at the
irradiation frequency. In such cases exces-

21.2 Survey of Characterization Techniques

sive amounts of secondary electron emission, local heating and direct chemical reaction are responsible for the damage.
One specific disadvantage of the photon
techniques however, is the fact that lateral
imaging capabilities are limited by the diffraction properties of the light. Thus, a
beam of light, using normal far-field optics
(the focusing optics located many wavelengths away from the sample), cannot be
focused to a point much smaller that the
wavelength of the light without the appearance of undesirable diffraction beams
at diverging angles. This provides no problem for X-rays since the wavelengths are in
the angstrom region anyway, but for visible and infrared radiation, the most valuable spectroscopic wavelength regions, the
beams effectively are limited to the |im
scale. In comparison, the best imaging capabilities of an ion beam (liquid metal) are
about 0.05 |im. Recent developments in Xray instrumentation allow comparable resolution for imaging of organic materials.
As a final general point, the in situ capabilities of photons can offer a distinct advantage for cases in which the analysis of
a surface is desired under conditions of
direct exposure to liquid or gaseous environments, e.g., electrochemical, biological,
or catalytic systems. Unlike particle beams
such as those of ions or electrons, which
effectively are scattered away with energy
dissipation upon entering such an environment, photons at all frequencies can be
transmitted long distances unless the media happen to be highly absorbing at the
frequencies of interest. In particular, Xrays and optical frequency beams can be
transmitted efficiently through most gases
and liquids including water, a particularly
important medium. Thus techniques that
operate at these frequencies such as
GIXRD, XRR, XRF, SE, SPR, Raman
and nonlinear spectroscopies can be used

677

to advantage for such in situ studies of


surfaces. On the other hand, a large number of liquids and gases have strong absorption bands in the infrared region, e.g.,
water, so in situ infrared spectroscopy is
limited to solvents or gases with minimal
absorption, to very limited frequency
ranges between absorption features or to
very short path lengths which minimize absorption.
For detailed presentations of the principles utilized in the above discussion the
reader can consult the following sources:
Born and Wolf (1992); Harrick (1967); Debe (1987); Jeh (1988); Parikh and Allara
(1992).
21.2.2.1 Grazing Incidence X-Ray
Diffraction (GIXRD)
X-ray diffraction has been one of the
most central techniques in structure analysis of materials over the years and has become a standard laboratory experiment.
However, its application to surfaces and
thin films requires significant effort as the
number of scattering centers drops dramatically compared to bulk experiments
and because the lattice spacings of interest
are usually parallel to the surface, i.e., they
are in-plane, which requires the X-ray
beam to skim the surface with an angle of
incidence extremely close to 90 (refer to
Fig. 21-2). These characteristics require
high quality X-ray optics and high intensity, well-collimated sources as well as very
smooth, fiat surfaces. As a consequence,
synchotron radiation is preferred although
in some cases, rotating anodes can be used.
Whereas application to inorganic materials has become well-established (Feidenhans'l, 1989), the application to organic
materials has lagged behind because of the
more demanding requirements for successful analyses compared to inorganic materi-

678

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

als. One major reason is that most organic


materials are comprised of elements in the
lower part of the periodic table and the
scattering cross sections for these are relatively low. Another reason is the low damage threshold encountered in X-ray irradiation of many organic materials. In spite
of these stringent requirements, examples
of the application of GIXRD to bulk organic materials, thin films and monolayers
are continuing to grow. It has been demonstrated that GIXRD can be a very useful
method for the determination of the
molecular orientation and crystal structure
in near-surface regions of bulk materials
(Buckley and Taylor, 1984). The major issue is how to obtain a surface selective
diffraction signal with minimal interference from the bulk patterns. In general, the
surface selectivity arises from the fact that
a beam of X-rays traversing a path of
length / through a material is attenuated
from its incident intensity Io to an intensity
/given by (Buckley and Taylor, 1984)
T T

l 70

(21-6)

where \i is the linear absorption coefficient


for the particular material concerned at the
appropriate X-ray wavelength. Further,
the effective depth yielding a diffraction
signal can be varied by varying the angle of
incidence. A particularly useful aspect of
X-rays is that over a range of X-ray frequencies, many materials exhibit a very
slightly lower real refractive index value
than that of air. Thus since air is the higher
index medium, a critical angle exists above
which value the reflection becomes total
with an evanescent wave penetrating into
the material, as depicted in Fig. 21-2 with
medium 2 as air. Thus for any value of the
angle beyond 0C, diffraction can occur only from a region governed by the effective
penetration depth (Eq. (21-5)). Factor
et al. (1991) have shown that for X-ray ra-

diation at the wavelength of 0.1558 nm,


the penetration depth dp(</>) for a polyimide polymer (see below) ranges from 5.0
to 10 nm for incidence angles past the critical angle 0C [refer to Eq. (21-4) and (21-5)
and Fig. 21-2].
A variety of sample types have been analyzed by GIXRD ranging from polymers
(Buckley and Taylor, 1984; Factor et al.,
1991) to Langmuir-Blodgett (LB), self-assembled monolayers and surfactant type
of films supported on solid substrates
(Amador et al., 1993; Pachence and Blasie,
1987; Fenter et al., 1991, 1993; Seul et al.,
1983) to surfactant and lipid films at the
air-water interface (Langmuir films) (Barton etal., 1988; Tippman-Krayer and
Mohwald, 1991; Wolf et al., 1988). In these
studies quite significant detail has been obtained with regard to the translational
symmetry and lattice spacings of these
films for even monolayer coverages. A
good example of a recent polymer study is
given by a report (Factor etal., 1991) of
the analysis of the near surface structure of
polymide [poly(pyromellitiodianhydride
oxydianiline); PMDA-ODA] in which it
was shown that diffraction patterns characteristic of crystalline PMDA-ODA
molecules are present under conditions for
near-surface region diffraction and that
this ordering extends ~9.0 nm into the
bulk of the specimen. In addition, the lateral registry of the molecules parallel to the
surface of the film extends over relatively
large distances of ~15.0nm. Thus the
analysis indicates that the surface regions
exhibit highly oriented and ordered polymer chains. As there are almost no other
ways to obtain such detailed information
of surface region ordering in such a direct
manner it can be expected that examples
will continue to grow. However, it should
be pointed out that while the technique has
been established as viable it is necessary to

21.2 Survey of Characterization Techniques

ensure that the samples to be studied are


very smooth and flat in order to achieve
successful results.
21.2.2.2 X-Ray Reflectivity (XRR)
In addition to the use of X-rays to generate diffraction patterns, simple reflectivity
measurements can be utilized for accurate
determination of physical characteristics
of parallel layer, thin film structures such
as thin polymer films on smooth substrates. These characteristics include layer
thicknesses, densities and interface roughness and are important in many studies
involving thin film samples such as polymers and liquid crystalline materials supported on smooth substrates.
The method depends upon very precise
measurements of the angular dependence
of the reflectivity of a collimated beam of
X-rays from a parallel-layer thin film sample. The underlying principle was discussed in the beginning of the main section
and essentially is embodied in Fig. 21-3.
Due to the refractive index (n) dependence
of the reflection-transmission ratios at
each phase boundary and the interference
effects caused by the presence of layers in
the sample, the overall sample reflectivity
depends upon the angle of incidence and
the X-ray wavelength in an exact way as
determined by the thicknesses of the layers
and their refractive indices. In the case of
X-rays, in contrast to optical and infrared
frequencies, the values of n in nonabsorbing regions are directly proportional
to the electron densities, a very useful simplification. Since the reflectivity is just one
known and there are several unknowns,
additional independent measurements are
made at a variety of angles. One can see
from Fig. 21-3 that in order to permit large
interference effects of the X-ray waves
within the layers it is useful to choose val-

679

ues of (j) which lead to path lengths within


the layers that are of the order of or longer
than the wavelength of the X-ray. For very
thin films, e.g., monolayers, and short
wavelength X-rays it is obvious that this
can require angles very close to grazing
and to the critical angle (at which the reflectivity becomes unity) which itself is
very close to grazing at X-ray frequencies.
This configuration in turn places a constraint on the sample that it be extremely
smooth and flat so that the beam can skim
across without scattering or without reflecting off at a variety of angles; although
small amounts of surface roughness of the
order of the wavelength of the X-ray or
less often can be tolerated. The experiments are best done on a synchotron in
order to achieve good signal to noise but
for high quality samples acceptable results
can be obtained with rotating anode
sources.
For optimal samples, the accuracies of
the measurements of layer thicknesses can
be excellent. It has been possible to determine the surface roughness, total film
thickness and interface width in polymer
thin films with ~ 1 nm resolution (Foster
et al., 1990). However, due to the relatively
small differences in electron densities between different polymers and typical support materials such as flat silica glass
plates, the reflectivity-angle curves show
poor sensitivity to changes in film structure. In order to use XRR effectively for
polymer interface analysis, one component
of the interface should contain heavier
atoms to impart sufficient electron density
contrast between the phases. For example,
Stamm et al. (1991) have used XRR to
study the interface between poly(methyl
methacryalate) and pol(vinyl chloride)
while Zhao et al. (1991) have investigated
the system of polystyrene-polybromostyrene. Wasserman et al. (1988) have used

680

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

XRR to measure the thickness and density


distributions of self-assembled monolayers
of alkylsiloxanes on silicon substrates and
extensions of these studies have been applied to characterization of adsorbed liquid overlayers (Tidswell et al., 1991). Figure 21-4 shows the XRR curves of alkylsiloxane monolayers with different chain
lengths on oxidized silicon substrates. Recently, Reiter et al. (1991,1992) have taken
advantage of the large difference in electron densities between metals and organic
materials to investigate the morphology of
the interface between thermally evaporated silver and Langmuir-Blodgett films.
For detailed discussions on the applications of XRR in polymer surface and interface analysis, the reader can refer to recent
publications by Stamm et al. (1991) and
Russell (1990).

10-

5-

en

0-

-5-

-10J
1.0

Figure 21-4. Intensity, R, of X-rays reflected from


alkylsiloxane monolayers on silicon-silicon dioxide
substrates as a function of qz, the momentum change
of the photon upon reflection. The monolayers were
prepared from alkyltrichlorosilanes, Cl3Si(CH2)nCH3.
The solid line is the calculated Fresnel reflectivity, R{,
for perfectly smooth silicon substrate. Reprinted with
permission from Wasserman et al. (1989). Copyright
American Chemical Society.

21.2.2.3 Spectroscopic Ellipsometry (SE)


In the XRR technique, as well as in
many other quantitative techniques using
light beams on transparent samples, the
precise value of the reflectivity of a thin,
parallel layer structure, measured under a
specific set of known conditions, allows
the computation of one sample property,
either a layer thickness or a layer refractive
index, as long as all the other properties
are known. If a film layer becomes absorbing then the refractive index has two
parts, real and imaginary (n and k, in
Eq. (21-1)) and the complete value of n
cannot be determined with one measurement. Ellipsometry is a technique which
has been developed to overcome this problem by providing to knowns in each measurement (Azzam and Bashara, 1977). The
extra information is derived by keeping
track of the polarization, viz, the direction,
of the electric field vector of the light, before and after reflection, as opposed to just
the intensity, the basis of the usual reflectivity measurement. The polarization
change can be measured by placing polarizers before and after the sample and the
changes in beam intensity can be measured
as a function of combinations of polarizer
angles using various strategies. If the ellipsometry measurement is combined with a
systematic variation of input wavelength
or angle of incidence, then a large number
of independent observations can be made
about a given sample and the opportunity
exists to analyze the sample structure in
detail, including such features as interface
roughness, void content and density gradients. Historically, ellipsometry has been
developed in the UV-visible region because high precision optical components
have been optimized at these wavelengths
and because a variety of solids possess important electronic spectra in this region.

21.2 Survey of Characterization Techniques

Because of the latter, the major choice has


been to vary wavelength systematically
and thereby obtain ellipsometric spectra
(Aspnes, 1976) but limited work has been
done with variable angles (Woollam and
Snyder, 1990; Debe and Field, 1991). In
recent developments, instruments have
been made which are capable of obtaining
complete, high precision spectra across
most of the visible region in fractions of a
second (Collins, 1990). Because of the
longer wavelengths of the light, UV-visible probes are inherently less sensitive to
the details of thin film structures compared
to X-ray probes, which have wavelengths
on the order of the critical film dimensions,
e.g., layer thickness and interface roughness, for nanometer-scale films. However,
the extremely high precision and obtaining
of two knowns in the UV-visible ellipsometric measurement compensates effectively for the longer wavelength relative to
a simple, less precise reflection measurement at short wavelengths.
Single wavelength ellipsometry has become a standard tool for determining the
thicknesses of organic films on smooth,
solid supports and numerous examples
exist for applications such as self-assembled monolayers, Langmuir-Blodgett films
and thin polymer films (Ulman, 1991).
However, for nonabsorbing, monolayer
thickness films, because of the nature of
the optical equations in this thickness
regime, both the film thickness and the real
refractive index cannot be determined
from a single wavelength measurement
(Aspnes, 1976). This problem is usually approached by assuming that the refractive
index of the film is the same as that of the
film material in bulk form (Allara and
Nuzzo, 1985; Porter etal., 1987; Wasserman etal., 1989), an assumption which
generally gives good results for high quality films. It is obvious that a simple advan-

681

tage of SE over a single wavelength measurement is that a number of independent


measurements are obtained which are redundant in the film thickness value to be
obtained. Thus highly improved statistics
can be obtained in the calculation (Collins
etal., 1993). The use of SE to analyze
phase segregation in thin films of polymer
mixtures has been reported (Sauer and
Walsh, 1991).
One useful advantage of ellipsometry in
the UV-visible region is that the measurement is virtually unaffected by the presence of a liquid phase surrounding the
sample, with the obvious caveat that the
liquid medium must have a different refractive index than the film of interest in
order that an optical interface exist. This
capability allows in situ measurements and
attempts have been made to follow adsorption processes of polymers this way.
Some workers have attempted to use in
situ single wavelength measurements to
provide independent values of thickness
and refractive index of adsorbing polymer
films with limited success (Kawaguchi and
Nagata, 1991) and attempts have been
made to measure concentration gradients
of polymers adsorbed at surfaces by utilizing evanescent wave probes generated
from prism configurations to localize the
analyses to the near-interface region (Kim
etal., 1989). Theoretical considerations
for analysis of such composition profiling
experiments have been reported (Charmet
and de Gennes, 1983).
The major advantage of SE arises when
the films possess nonzero values of the
imaginary optical function (traditionally,
dielectric functions rather than refractive
indices are used for ellipsometry) and thus
give rise to spectra. One particularly useful
case arises when the optical functions of
the material are already known from independent measurement and the SE spectra

682

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

can be used for accurate determination of


film thickness and microstructure. This approach has been utilized in the electrochemical polymerization of pyrrole on
gold electrodes with fast SE measurements
to determine the real time evolution of the
nucleation and growth of the films with
time resolution of 50 ms (Kim et al.,
1991). This study is a good example of
both the in situ and time-resolved capabilities of SE. Based on this kind of example
it appears that there are many opportunities for the applications of SE, especially
time-resolved SE, to a variety of problems
involving quantitative characterization of
the physical structure of thin organic films.
21.2.2.4 Surface Plasmon Resonance
(SPR)
Surface plasmon resonance is a highly
specific optical effect which involves the
coupling of a beam of light into a thin
metal film under a very narrow range of
experimental conditions and is manifested
as a sharp change in reflectivity. In order
to give rise to the effect the metal must
have optical functions (Eqs. (21-1) and
(21-2)) within a narrow range of values for
the frequency of the light to be used. The
coupling of the light into the film can be
accomplished in several ways but the most
common methods involve the use of contiguously placed prisms, of suitable refractive index values, which accomplishes coupling via evanescent waves (Fig. 21-2) and
the use of metal films configured as gratings which allows coupling via grating
modes. However, for a thin film analysis
only the prism-based coupling schemes are
convenient. The simplest way to generate a
surface plasmon this way is to propagate
a light beam of the correct frequency
through a transparent, relatively high refractive index prism coated on the back

side with a thin ( 50 nm) film of an appropriate metal, such that that the light
will reflect off the back, inside surface (the
metal-prism interface) of the prims at a
controlled angle. As this angle is varied,
the reflectivity slowly changes until a narrow range of angles is reached when a
sharp dip in the reflectivity occurs. Past
this range, the reflectivity then returns to
higher values. The dip occurs because the
electrons of the metal undergo a collective
oscillation with the light, a response
termed a surface plasmon polariton mode
or a surface plasmon resonance, which requires that the component of the propagation vector of the light parallel to the interface have a very specific value as determined by the optical properties of the
prism and the metal film and by the frequency of the light. The magnitude of the
wavevector component is tuned by adjusting the incidence angle. The value of the
experiment is that the exact angle for the
plasmon resonance peak is extremely sensitive to any overlayer on the metal film,
even monolayer to submonolayer thicknesses. Based on this dependence, fashioned upon classical electromagnetic theory, the peak position can be used to determine the properties of an attached film, in
particular the film morphology, thickness
and density, similar to an ellipsometry
analysis. Further, the electric field which
exists at the surface of the metal is a decaying evanescent field and so the SPR effect
is surface selective as well as surface sensitive, with a range scaled according to the
wavelength of the light, similar to the usual
penetration depth of evanescent fields,
lOOnm or more for visible light. Because visible light is involved, the experiment can be conducted in a variety of nonabsorbing liquids, such as water, and thus
in situ analyses are possible. On this basis
the SPR method has found utility as a

21.2 Survey of Characterization Techniques

technique for monitoring the adsorption


of molecules and polymers at metal surfaces almost exclusively Ag or Au (Pollard
and Sambles etal., 1987; Zhang etal.,
1987; Herminghaus and Leiderer, 1989;
Huang et al., 1989; Mayo and Hallock,
1989; deBruijn et al., 1991). Recent developments have included the use of imaging
optics to provide spatial resolution of
~10|im (Kooyman and Krul, 1991).
More recently, the SPR effect has been
used as a detection method for biomolecules (Lofas et al., 1991). One distinct
advantage of SPR analysis is the fact that
quite good time resolution is possible since
the surface plasmon can be generated and
measured in quite a short time. Real time
analyses can be made from a scale of seconds (Mayo and Hallock, 1989), useful,
for example, for macromolecular absorption, to a scale of nanoseconds (Herminghaus and Leiderer, 1991), useful, for example, for photo-induced relaxation processes in thin films. Another distinct advantage
of SPR is the fact that optical wavelength
light is generally unaffected by solvents
and thus in situ analyses are possible.
From this point of view one should expect
an increase in applications to the study of
in situ, real-time adsorption of molecular
and polymer films for cases where thin noble metal films are suitable substrates.
21.2.2.5 Raman Spectroscopy and SurfaceEnhanced Raman Spectroscopy (SERS)

Standard Raman Spectroscopy


Raman spectroscopy is a scattering experiment in which a photon of one frequency scatters from the target solid and
produces output photons with frequencies
shifted either positively or negatively from
the input values by exactly the equivalent
energies for transitions between specific
energy levels of the target material. Figure

683

21-1 depicts the basic experiment. In standard Raman scattering the scattered light
exists over a wide range of angles (0OS in
Fig. 21-1). Thus, to maximize the signal,
the light must be collected over a wide
viewing angle and focused efficiently into
the detection system. The Raman effect occurs optimally in the range of ultraviolet
through visible frequencies and is valuable
because the frequency shifts of the light
primarily arise from excitations of vibrational levels in the scattering material and
the energies of these levels are quite sensitive to chemical bonding structure. In this
regard Raman spectra would appear to
give the same information as infrared spectra. However, Raman spectra are not
equivalent to infrared spectra because of
differences in the spectroscopic selection
rules which govern the probability of a
particular energy level transition appearing in the spectrum. The most significant
generalization which can be made in this
regard is that there are generally fewer Raman modes to be observed than for infrared spectra and that the Raman modes
arise primarily from the ordering and symmetry characteristics of the material, in
contrast to infrared spectra which are generally less sensitive to such features. Thus
Raman spectra are useful complements to
infrared spectra, and in fact, and in many
cases of ordered materials would be preferable. However, in the application to surface and thin film analysis several serious
problems arise which have prevented the
widespread use of Raman in this area. The
major problem is the extremely low scattering cross section, typically only one out
of ~10 6 photons produces a Raman output with the remainder scattering off at the
input frequency (Rayleigh scattering). This
characteristic leads to extremely low surface sensitivity and only can be overcome
by extremely low noise detection of the

684

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

scattered photons and very high input intensity from a laser source. However, the
latter has limits before sample damage occurs and thus the useful advances have
been made by optimizing the detection system. Several groups have reported monolayer sensitivity for organic molecules by
utilizing very advanced, multichannel detection systems (Dierker et al., 1987; Perry
and Campion, 1991; Pemberton et al.,
1990; Bryant and Pemberton, 1991 a). Further, many materials exhibit intrinsic fluorescence background emission which interferes with the Raman signal detection and
thus only a limited number of materials
can be usefully analyzed for low levels of
scatterers, such as the monolayer-scale experiments above. However, as the film
thickness increases to tens of nanometers
or more, one can expect increasing success
in obtaining Raman spectra of the entire
film. In this regard, the other major problem is the lack of surface selectivity, as is
typical for typical photon in-photon out
experiments. Thus excitation of a typical
solid polymer, for example, would produce
a signal coming from deep within the material and, in view of the low surface sensitivity, with no discernable surface layer
contribution. To overcome such problems
of surface selectivity, one can resort to
containing the exciting electric field at a
boundary by using internally reflected
light in a prism to create an evanescent
wave as depicted in Fig. 21-2 b and c. Raman spectra of thin polymer films have
been reported for slab waveguide geometries with the polymer films coated on the
waveguide surfaces (Rabolt etal., 1981;
Zimba et al., 1990; Miller et al., 1987). In
general, spectra can be obtained in this
way for film thicknesses of tens of nanometers and for thicker films the spectra will be
confined by the evanescent wave localization to depths of fractions of the wave-

length of the exciting light. The varieties of


similar approaches which can be made in
obtaining thin film Raman spectra has
been reviewed (Bohn and Walls, 1991).
Because Raman spectroscopy almost always involves visible light, it offers the potential advantage of in situ analyses, since
most liquids are highly transmitting at
these frequencies. Such an application has
been reported for an electrochemical environment (Shannon and Campion, 1988).
The use of visible light also means that
lateral imaging capabilities should exist at
resolutions up to ~0.5 (im, the diffraction
limit of the input light beam. Raman microscopes are commercially available but
generally the highly converging optics are
not compatible with strategies to obtain
either high surface sensitivity, particularly
monolayer level, or surface selectivity with
evanescent wave geometries.
Surface-Enhanced Raman Spectroscopy
(SERS)
The standard optics-based strategies discussed above for improving the surface
sensitivity and selectivity of Raman scattering have not yet provided the level of
performance generally desired for a typically useful surface and thin film analysis
tool. However, recent developments in the
dependence of the scattering intensity of a
supported thin film as a function of the
specific electronic structure and surface
morphology of the substrate have produced enormous increases in both surface
sensitivity and selectivity. These developments have led to the a sample-based strategy termed surface-enhanced Raman scattering (SERS). With this approach, Raman scattering cross-sections of molecular
species adsorbed onto roughened surfaces
of certain highly conductive metals have
been enhanced by as much as 106 com-

21.2 Survey of Characterization Techniques

pared to normal Raman scattering crosssections (Allara et al., 1983; Boerio et al.,
1991). Enhanced Raman scattering of
molecular species has been observed on a
variety of metal substrates including Au,
Cu, Ag, Al and Li. It has been found that
SERS is quite surface selective as shown by
observations that signal is obtained only
from those molecules that are within a few
monolayers away from the metal substrates. However, it should be realized that
normal Raman spectra will still occur so
that as the total thickness of a film increases to values on the scale of larger than
~100nm, the normal Raman scattering
signal from the entire film also will began
to be observed. Within this constraint
however, SERS is considered as highly surface and interface sensitive (Allara et al.,
1983; Venkatachalam etal., 1988) and
should be an ideal technique for the analysis of the adsorption of organic species on
appropriate metal surfaces (Bukowska
et al., 1991). SERS has been used by many
researchers in the investigation of a variety
of problems including corrosion inhibition
of metal surfaces (Bukowska etal., 1991;
Xue and Zhang, 1991), metal-polymer
adhesion (Boerio etal., 1991; Tsai etal.,
1991 a) and orientation and conformation
of molecules on metal substrates (Xue
etal., 1991; Tsai etal., 1991b; Tsai etal.,
1992). Recently SERS has been used in the
investigation of the interdiffusion at polymer-polymer interface (Hong et al., 1991)
and in the study of self-assmbled monolayers on metal substrates (Bryant and Pemberton, 1991). One of the advantages of
SERS over other surface sensitive techniques is that no vacuum is required for the
analysis and in situ analyses in a variety of
liquid environments is possible. One of the
major drawbacks of SERS is in fact, rooted in its basis, the need for highly specific
substrates. The only choices which seem

685

usable are the coinage metals, Cu, Ag and


Au, with highly roughened surfaces or in
the form of clusters deposited on a smooth
surface such as glass. In both cases the
typical length scales of the morphological
features are of the order of ~ 5 - 5 0 nm.
This roughness generally will preclude
quantitative work and could interfere with
the ordering and organization of overlayer
films compared to smooth substrates
(Knoll et al., 1982). On this basis, it would
appear that SERS is to be considered as a
surface analysis tool of only occasional use
when the sample of interest meets the stringent SERS requirements.
21.2.2.6 Nonlinear Techniques: Second
Harmonic Gain (SHG) and Sum Frequency
Generation (SFG)

Within the past decade or so a number


of interesting and quite useful optical effects have been discovered which derive
their basis from a nonlinear dependence on
the intensity of the incoming light beam
(Shen, 1984). As a consequence of the
higher order dependence, laser irradiation
is necessary to supply sufficiently high intensity to be seen relative to the usual linear processes which are the basis of most
common spectroscopies and optical effects. A variety of attempts have been
made to adapt these nonlinear techniques
to the analysis of the organic materials at
surfaces and interfaces but by far the only
reasonable applications appear to be associated with second harmonic generation
(SHG) and infrared-visible sum frequency generation (SFG). For this reason the
discussion below will be restricted to these
two techniques. The principles and applications of SHG and SFG have been well
documented recently (Anderson, 1993;
Eisenthal, 1992; Vogel and Shen, 1991;
Corn, 1991; McGlip, 1990; Shen, 1989;

686

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

Richmond et al., 1988; Chemla and Zyss,


1987) and the interested reader is referred
to these publications for details. Both techniques offer significant advantages in interfacial characterization of organic materials in that the analyses can be nondestructive, interface selective, submonolayer sensitive, in situ with both gases and
liquids and capable of time-resolution on
the nanosecond to subnanosecond time
scale. The SHG experiment is far easier to
run than the SFG but offers significantly
less information about the molecular
structure of the material analyzed.
In an SHG experiment, a laser beam at
a selected frequency is directed at a flat
surface and, at a sufficiently high input
power, a measureable fraction of the beam
is emitted as light with a doubled frequency. The doubled beam is isolated from the
larger intensity, reflected beam at the original frequency by the use of a filter or
monochromator and then the intensity of
the doubled beam is measured. In addition, the effects of electric field polarization of the input and output light beams
can be monitored. At a given input intensity, the intensity of the output doubled
beam is a function of the electronic structure of the constituent molecular groups
responsible for the optical response. In this
regard there are three particular features of
the material which are critical to the utility
of the experiment. First, if the electronic
structure of the molecular groups is anisotropic then the SHG response will depend
upon the orientation of the groups with
respect to the sample surface. This effect
can be taken advantage of to determine the
molecular orientation of interfacial molecular groups, particularly for strongly responding, anisotropic chromophores such
as - N O 2 and - C = N substituted aromatic
compounds. Second, because of the intrinsic nature of the second-order optical

response with respect to the intensity of


the incoming electromagnetic field, only
groups which are organized in a fashion
which provides a directionally nonequivalent surrounding around each group, viz, a
noncentrosymmetric arrangement (the environment at location x away from the
group is different than that at -x), will
induce the SHG effect. This structural
requirement is satisfied, of course, in
noncentrosymmetric crystals but, most importantly, is satisfied at an interface, which
by definition must have a different environment on each side of the interfacial
groups. The interfacial selectivity holds
even if a thick overlayer film is present
above the interfacial layer as long as the
overlayer film does not have a noncentrosymmetric organization and as long as
it is relatively transparent to the incoming
and outgoing light. Thus SHG is one of the
very few general interface-selective characterization techniques. It follows from this
consideration that the presence of a nonabsorbing liquid phase can be straightforwardly accommodated to allow in situ
measurements and this has allowed studies
of interfacial molecular structure (Eisenthal, 1992) including the gas-liquid interface, for example in Langmuir films (Vogel
and Shen, 1991), and the liquid-solid interface, particularly in electrochemical systems (Corn, 1991; Richmond etal., 1988;
Campbell et al., 1990). Third, the SHG response is proportional to the number of
responsive molecular groups present at a
surface or interface and thus quantitative
measurements of coverage can be made,
for example at the electrochemical interface (Corn, 1991). In this regard, the technique often is capable of submonolayer
sensitivity as long as the intrinsic secondorder optical response is sufficiently high,
a condition usually easily met for aromatic
and conjugated groups with polarizable

21.2 Survey of Characterization Techniques

electron clouds. Recent applications of


SHG to interfacial analyses have included
thermal relaxation effects in multilayer LB
films of polymers (Senoh et al., 1990).
In the SFG experiment, a fixed-frequency, visible laser beam and a tunable, infrared laser beam are superimposed at an interface whereupon, under the correct conditions, light is emitted at a frequency
which is the sum of the two incident frequencies. The intensity of this sum frequency beam is measured and when the
frequency of the tuned infrared beam
matches an infrared and Raman active
transition frequency of a vibrational mode
in the irradiated material (resonance condition), a change in the intensity of the sum
frequency signal is observed. Thus the
SFG experiment can provide a vibrational
spectrum of the interacting material and in
this regard is much more useful for characterization than SHG. Howeer, the SFG experiment requires considerable intensity of
the input laser beams and thus pulsed
beams with suitable peak powers are utilized, a more difficult requirement than for
SHG. Since the same overall physical principles apply to both the SFG and SHG
processes, it accordingly follows that SFG
is interface selective. The major applications to date have been in organic monolayers (Eisenthal, 1992; Vogel and Shen,
1991) and spectra have been observed both
for self-assembled monolayers on gold and
LB films (Harris etal., 1987; GuyotSionnest etal., 1987). Ong etal. (1992)
have carried out in situ studies of the solvent-monolayer interface in self-assembled monolayers using SFG. These spectra
are typical of the reasonably good signalto-noise ratios which can be obtained and
indicate that submonolayer sensitivity is
quite possible for organic films, particularly since a number of instrumental improvements appear possible. In addition, studies

687

of vibrational state relaxation dynamics in


self-assembled monolayers have shown
that time resolution into the picosecond
scale is possible (Harris, 1991). Bain etal.
(1991) have demonstrated that the intensity of an SFG spectral response is directly
proportional to the number of molecules
at an interface and thus can be used for
quantitative analysis of absorbed molecular species. Applications of SFG to surface
analysis problems now seem possible as
illustrated by a recent study of diffusion
and cluster formation in the deposition of
metal films on a polyimide surface (Zhang
etal., 1992).
21.2.2.7 X-Ray Fluorescence (XRF)
and UV-Visible Fluorescence

Fluorescence can be viewed as a scattering experiment, illustrated in Fig. 21-1, in


which an input photon is converted to an
output photon of a different frequency,
usually of a lower value. The output radiation exits over a wide range of angles, and
like Raman scattering, is collected and focused into a detection system. However,
unlike Raman scattering, the signals from
fluorescence are usually quite strong and
the technique can exhibit good surface sensitivity. Fluorescence arises when the input
radiation creates an excited state in the
target material and this excited state subsequently radiatively decays back to the
ground state with emission of radiation of
similar frequency to the excitation frequency. Since a variety of excitations are
possible, from atomic core-level to vibrational, the experiment can be carried out
from X-ray to infrared frequencies. The
fluorescing signal can be utilized as a structure probe by interpretation of the spectral
details in terms of the chemical environment and/or the concentration and location of the excitation center. For surface

688

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

and thin film analysis the X-ray and U V visible regions have been useful. However,
a restriction on the experiment is that excitation frequencies must match allowed
transitions in the material to be studied
and further, not all excitations efficiently
fluoresce. This condition is actually rather
restrictive since not many organic materials will fluoresce at convenient wavelengths with acceptable efficiencies. For
this reason, most fluorescence experiments
involve the use of highly efficient fluorescening tag groups or atoms which are
placed in the material of interest to provide
probes of the surrounding structure or to
locate the position of the probe with respect to the surface and other tag groups.
In the X-ray region tag groups are often
heavy atoms whereas in the optical region
organic chromophores are used, such as
aromatic ring compounds. As with most
photon emission techniques, there is no
general surface selectivity so the experiment usually consists of analyzing thin
films of material on a substrate. This mode
works quite well in the UV-visible frequency region where highly efficient fluorescing tag groups can be placed on polymer chains or diluted into molecular films
to provide a probe which senses the local
environment through matrix-induced
shifts in the fluorescence characteristics of
frequency, linewidth and intensity (Chen
and Frank, 1989). However, it should be
noted that the introduction of the tag
group may cause an undesirable perturbation in the structure under study and thus
fluorescence tagging experiments must be
carefully evaluated for each application.
Of much more specific interest to surface
analysis have been the use of optical configurations to produce evanescent waves
which can localize the excitation electric
field at an interface (see Fig. 21-2). In the
X-ray region these experiments can take

advantage of the fact that the real refractive index (n in Eq. (21-1)) of many materials is less than one and thereby provide a
means to generate an evanescent wave inside a medium via external reflection of an
X-ray beam from the ambient side at an
angle above the critical value (see also
Sees. 21.2.2.1 and 21.2.2.2). Bloch et al.
(1985) first utilized this configuration at
the air-water interface to depth profile the
near-surface concentration gradient of a
dissolved polymer by measuring the fluorescence yield of a heavy atom tag on a
polymer. Because the penetration depth is
so shallow at X-ray wavelengths (values of
~ 5 nm or less; see Eq. (21-5)) the depth
resolution in evanescent XRF experiments
can be in the nanometer range. In another
variation, a brominated solvent was used
and the bromine X-ray fluorescence was
detected to provide a density profile (Barton etal., 1992). Similar experiments can
be carried out in the UV-visible region
using fiuorescing chromophore groups
sensitive in this region and changing the
optical configuration to a high refractive
index prism which generates an evanescent
field in a solution adjacent to the back face
of the prism, viz, the side opposite to the
one irradiated by the excitation light (Allain et al., 1981). The use of both the optical and X-ray wavelength evanescent wave
fluorescence probes for depth profiling
concentrations has been reviewed (Rondelez etal., 1987). A general review of
depth profiling including optical fluorescence techniques has been published (Bohn
and Miller, 1991). The possibility of high
spatial resolution for XRF of monolayers
appears likely in view of recent advances in
X-ray instrumentation which allows resolution on the ~ 50 nm scale (Da Silva,
1992; Meyer-Use etal., 1991).

21.2 Survey of Characterization Techniques

21.2.2.8 Infrared Spectroscopy (IRS)


The physical mechanism of IRS is well
understood in terms of the absorption of a
photon of infrared frequency with excitation of a given vibrational mode from a
lower energy (almost always the ground
state) to one of higher energy. Alternatively, for emission, a photon is emitted by
relaxation of an excited vibrational state
to a lower one. Because the vibrational
modes of molecules are extremely sensitive
to changes in chemical bonding and molecular symmetry, IRS has been a traditional
tool for characterization of organic and
organo-metallic molecules (Nakamoto,
1986; Lin-Vien et al., 1992; Bellamy, 1975).
Using appropriate sample structures and
experimental configurations, a variety of
extensions of IRS have been made to surface and thin film analysis and have proven to be very useful in characterizing subtle
aspects of structure (Yates and Madey,
1987; Chabal, 1988; Allara, 1993; Ishida,
1987).
The simplest absorption experiment is
transmission through a thin film sample.
More complicated absorption experiments
involve specular reflection from external
and internal surfaces of flat, smooth samples. In external reflection the beam is reflected from the front (outer) surface of a
flat sample, either a bulk material such as
a polymer or a substrate with a film (or
stack of films) of interest on it, e.g., a surfactant film on a metal sheet. The external
reflection configuration essentially is given
in Fig. 21-2 a in which a film would be
placed in medium 1 at the interface between 1 and 2. In internal reflection, a
beam of light is transmitted into a high
refractive index, transparent (nonabsorbing) prism, e.g., silicon, and reflected
internally off a face of the element containing a film (or stack of films) of interest on

689

the opposite (ambient side) (Harrick,


1967). The internal reflection configuration is shown in Fig. 21-5. For all of the
above absorption experiments one requires a source of IR radiation of variable
frequencies, a radiation detector and a
mechanism for separating the response of
the sample for each frequency interval of
light (monochromatic response). There are
a large number of choices for these components, for example, a synchotron can
provide an excellent source of broadband,
collimated far IR radiation. By far, the
most common instrument for the majority
of surface-thin film analyses which involve mid-IR frequencies is a Fourier
transform infrared (FTIR) spectrometer
(Griffiths and de Haseth, 1986). A general
point to be made with regard to instrumentation is that typically good FTIR systems
operating in the mid-IR under medium
resolution conditions can provide >0.1 %

SINGLE REFLECTION

OVERLAYER
FILM

MULTIPLE REFLECTION

Figure 21-5. Schematic representation of the sample


geometry in internal reflection measurements.

690

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

precision in intensity per scan. The signal


to noise can be increased by up to nearly
two orders of magnitude by signal averaging. This level is more than sufficient to
detect monolayer quantities of a film, even
for non-optical sampling conditions. However, one should note that other specific
combinations of components may solve
critical surface or thin film analyses problems in some cases better than typical commercially available instruments.
As with all first order or linear optical
experiments (see beginning of Sec. 21.2),
surface selectivity only arises when the
light is guided along a material phase
boundary to create an evanescent wave
(see Fig. 21-2 b and c) and under these conditions the penetration depth (dp in
Eq. (21-5)) is restricted just to a value of
the order of a micrometer for the mid-infrared region. For this reason IRS,
whether internal or external reflection or
transmission, is only useful for real surface
analysis in the cases of samples in which
the surface signal, defined to be associated
with a region the order of several monolayers or less, can be distinguished easily from
a much larger total signal. Such samples
include, for example, ultrathin polymer or
molecular films supported on a spectrally
featureless substrate or a bulk polymer
possessing a modified surface layer with a
spectrum completely different from the
bulk.
The strategies for obtaining spectra
from the above thin films vary with the
exact nature of the sample. Consider the
case of a thin film supported on a smooth,
planar substrate. If the substrate were a
highly infrared-absorbing (opaque) material such as the reflective metals Au and Al,
external reflection would be the only simple method appropriate. External reflection from metals is probably the most
common example of IRS thin film analysis

(Porter, 1988) and has been utilized for a


variety of studies including, for example,
self-assembled monolayers (Allara and
Nuzzo, 1985; Labinis et al., 1991; Dubois
and Nuzzo, 1992; Walczak etal., 1991),
Langmuir-Blodgett films (Allara and
Swalen, 1982; Umemura etal., 1990;
Rabolt et al., 1983) and thin polymer films
(Lenk et al., 1993; Kelley et al., 1987). An
example of an external reflection spectrum
is given in Fig. 21-6 where the C - H
stretching modes of a self-assembled
monolayer of an alkanethiolate on a gold
surface are shown (Parikh and Allara,
1992). Methods have been developed recently for precise quantitation of spectra
(Parikh and Allara, 1992). In particular,
the latter work has shown that for the case
of a planar, parallel layer sample, the spectra can be understood quantitatively in
terms of layer thickness(es) and orientational ordering of the component molecular groups provided that values of the frequency-dependent optical functions, n(co),
of the constituent layers are known. The
most general case of oriented films (anisotropic) requires the use of directionally dependent optical functions, viz., tensor
rather than scalar representations. An example of the application of the above theory can be seen in Fig. 21-6 which shows the
simulated spectrum of the alkane thiolate
monolayer in which the simulation was
based on a model structure of pairs of alltrans chains, with one chain twisted roughly 90 to the other, both chains tilted at 26
from the surface normal and a small fraction of end gauche defects in a large ensemble of these pairs (Allara and Parikh, 1992;
Laibinis etal., 1991). Finally, it is important to note that the external reflection
technique does not allow depth profiling
since the electromagnetic fields generated
at the surface extend as far away from the
surface as the width of the incoming beam

21.2 Survey of Characterization Techniques

0.006 -

691

C 18 H 37 S/Au

gauche allowed two chain model

2918

experimental data

0.003 0.002 -

- 2878

cc

0.004 -

- 2850

-log( Ho)

0.005 -

0.001 -

Figure 21-6. External reflection spectrum showing


the C - H stretching modes
of a self-assembled monolayer of an alkanethiolate
on a gold surface.

\n\

\}\

0.000 -

2700

ID
<O
0)
(M

2800

2900

3000

31

wavenumbers (cm

(usually several millimeters) and are far


too slowly varying as a function of distance to allow interpretation of the spectra
in terms of vertical mapping of signal contributions at any useful distance resolution. It is also possible to obtain emission
spectra of monolayer films on reflective
metals (Allara et al., 1984; Chiang et al.,
1987), but the experiment requires a specialized, cryogenically cooled spectrometer
to achieve sufficient signal-to-noise ratios
and thus is not a general method. On the
other hand, when the substrate is transparent to infrared radiation, e.g., Si, then
transmission or external reflection can be
used and when the substrate is a high refractive index material with a prism-like
shape, internal reflection also can be used.
In the past, transmission was used very
little because the spectra are quite weak for
monolayer coverage films of typical organic materials. However, recent advances in
FTIR instrumentation now have made
such analyses feasible (Sheen and Allara,
1993; Chen and Frank, 1989) because of
improved signal-to-noise ratios. External
reflection with semitransparent and transparent substrates presents a similar situa-

tion with regard to signal-to-noise ratios,


but, in addition, the fundamental nature of
the interaction of the electromagnetic
fields with the sample gives rise to considerably more complex spectra than obtained from transmission or from reflection with metal subtrates and the additional complexity provides severe difficulties
with quantitative analysis of film structure
and thickness. For this reason there are
few examples of detailed structural analyses for these types of samples, but recent
theoretical developments, discussed briefly
above for the case of metal substrates, now
allow quantitative interpretation of the
complex spectra associated with weak
metals and nonmetals (Parikh and Allara,
1992). An example of an external reflection
spectrum from a nonmetal is given in
Fig. 21-7 which shows in C - H stretching
mode region of a Langmuir-Blodgett
multilayer stack (5 layers) supported on a
carbon substrate (Parikh et al., 1993). The
figure also shows a simulated spectrum,
based on the above theory (Parikh and Allara, 1992), for a model structure of a single average all-trans chain tilted at 8 from
the surface normal. External reflection

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

692

(C 1Q hL Q C00) 9 Cd/ C
CO

cc
O

calculations
experiments
-6

2700

2800
2900
3000
Frequency (cm )

3100

Figure 21-7. External reflection spectrum showing


the C - H stretching mode region of a LangmuirBlodgett multilayer stack (5 layers) supported on a
carbon substrate.

IRS spectra have been reported recently


for self-assembled monolayers on materials such as semiconductors (Sheen et al.,
1993; Krishna etal., 1992).
In principle, the same signal-to-noise
problems exist with internal reflection as
with transmission, as mentioned above.
However, this problem was overcome a
number of years ago by using a long, thin
element or prism with parallel sides capable of allowing the beam, when passing
through the prism in the long direction, to
bounce off the internal surfaces a number
of times before exiting out the end as
shown in Fig. 21-5 (Harrick, 1967). These
elements are fabricated by cutting an infrared transparent material into a slab (trapezoid or parallelogram) or cylindrical rod
(conical ends) which is then polished to
give very smooth surfaces. Typical numbers of these multiple internal reflections
are 10-50 and the transmission efficiency
can be made to be acceptably efficient by
adjusting the input angle to the prism such
that the reflections occur at values above
the critical angle thus guaranteeing total

internal reflection and no transmission


leakage out the sides. Since each reflection
provides an additional contribution to the
total signal, the spectral intensities can be
increased over a single reflection by at least
an order of magnitude. Internal reflection
prisms work best when the refractive index
is high and optimal materials include Si,
Ge, KRS5 and ZnSe. Many examples exist
for the analyses of adsorbed films and
include Langmuir-Blodgett monolayers
(Kimura etal., 1986; Ahn and Franses,
1992), self-assembled monolayers (McGarvey etal. 1991; Tilman etal., 1989)
and thin polymer films (Culler et al., 1983;
Ancelein etal., 1990; Johnson and Granick, 1992). One obvious disadvantage of
internal reflection is the fact that many
substrates of interest, with respect to their
relevance in bonding or adsorbing organic
films, are not suitable for utilization as internal reflection elements.
A major advantage of the internal reflection experiment is that under conditions of total internal reflection, the only
light which appears outside of the sides of
the prism is in the form of a decaying
evanescent wave. This offers three advantages. First, if a sample of polymer is
pressed in firm contact with a side face of
the prism, the evanescent wave will penetrate a micrometer or so, according to
Eq. (21-4), into the surface region and infrared spectra can be obtained. However,
one should note that when the contact of
the film with the element surface is incomplete, errors in quantitative analyses accordingly arise. Although this method certainly is not a surface analysis in the typical sense of a few layers of material, it does
offer a way of limiting the depth region of
the surface studied, a decided advantage
over a transmission or external reflection
experiment which both sample the entire
bulk of the material. This application is

21.2 Survey of Characterization Techniques

used extensively to characterize the deep


surface regions of polymers modified by
various treatments (Bee and McCarthy,
1992). Further, since the direction of the
electric field is precisely defined at the surface of the prism when the input beam is
polarized, the spectra can be used, when
the directions of the vibrational motions of
the atoms and their associated dipoles are
known, to determine the orientation of
molecular groups in the overlayer surface
region. This approach has been used to
analyze for orientation effects in polymer
surfaces (Hobbs et al., 1983; Mirabella,
1988; Walls, 1991). A second advantage is
that in situ adsorption studies can be carried out in liquid (or gas) media much better than can be done by external reflection
or transmission. In these experiments, one
face of the prism can be exposed to the
medium and the spectral influence will be
limited to the effective penetration depth,
thus minimizing spectral contributions
from the medium. This approach has been
increasingly used for polymer adsorption
studies (Frantz and Granick, 1991; Johnson and Granick, 1992). It should be
pointed out though that external reflection
IRS has been used under special circumstances for analysis of electrode surfaces in
an electrochemical cell (Stole et al., 1991;
Korzeniewski and Pons, 1987). The third
advantage of the evanescent wave in internal reflection is that depth profiling of attached polymer films is possible by carefully utilizing the function form of the decaying electric field in the film overlayer. It
can be seen in Eq. (21-4) that the decay
function can be varied by changing element materials and angle of incidence for
appropriate types of prisms, which allows
generation of a set of independent spectra
each dependent upon concentration gradients in the film interior in a measureably
different way (Bohn and Miller, 1991). So-

693

lution of the equations allows, in good cases, depth resolutions of better than tens of
nanometers. In general, however, such applications are rarely feasible for practical
problems. The most common practical
problem which arises in depth profiling a
polymer surface region is the lack of completely reproducible, uniform elementpolymer contact from element to element.
This problem precludes quantitative profiling analyses although qualitative analyses should be successful.
Limited lateral imaging is possible with
IRS. Quite high quality IR spectroscopic
microscopes are commercially available
and provide spatial resolution with optimal measurements to near the diffraction
limit, > 10 jim (Messerschmidt and Harthcock, 1988). Imaging is best done when the
angle of incidence is near normal to the
surface and when the dependence of the
spectra on angle of incidence is not strong,
a factor which arises since high focussing
requires highly conical beam shapes. This
is suitable for transmission which uses normal incidence and for which the angle dependence of the spectra is not strong.
However, since reflection experiments with
thin films usually require nonnormal angles of incidence ranging from grazing to
~60, only elliptically shaped beam spots
with serious beam divergence can result
from focussing on a tilted surface and thus
lateral resolution is degraded, particularly
for grazing angle values of 80-85 typically used for external reflection on metal
substrates. Further, the angle dependence
can be quite strong which leads to quite
serious errors in quantitation of spectra
unless the precise spreads of angles and
accompanying intensities are known.
In cases for which surface films are to be
analyzed on powders or high surface area
granular materials the techniques of diffuse reflectance and photoacoustic spec-

694

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

troscopy are quite applicable. Little will be


said about such analysis here since these
techniques are well documented for powders (Griffiths and DeHaseth, 1986;
Mackenzie, 1988; Leyden and Murthy,
1988; Belton etal., 1987, 1988) and since
quantitation of surface coverage, derivation of detailed structure such as molecular
and bond orientation, and imaging are
rendered extremely difficult because of the
ill-defined analytical nature of the lightsurface geometry and the distribution of
sample characteristics such as surface morphology. However, two comments will be
made. First, qualitative analyses of overlayer films, even monolayers, are quite
simple, can be extremely useful and can be
carried out with commercial instrumentation. Second, whereas diffuse reflectance
works only for diffusely scattering samples, photoacoustic spectra can be obtained, even for monolayers on highly
smooth, planar substrates by attaching a
piezoelectric element directly to the substrate in such a fashion as to very sensitively detect the acoustic pulses generated
from absorption events in the monolayer
during infrared irradiation (Rothberg
et al., 1987). Such an experiment, however,
is sufficiently difficult such that it should
be considered a research tool rarely suitable for practical purposes.
Emission spectroscopy is a potentially
useful and underutilized mode of obtaining infrared spectra of surface films, which
are considerably thicker than monolayer
coverage (Griffiths and DeHaseth, 1986;
Mackenzie, 1988; Chabal, 1988; Compton
et al., 1991). All the above modes have referred to experiments in which spectra
were obtained by measuring the fraction of
light beam power absorbed by various infrared active groups in the irradiated sample. However, every sample always emits a
very weak but continuous infrared spec-

trum from the decay of thermally populated excited vibrational quantum states and
this spectrum contains contributions from
the same individual groups as seen in absorption. Thus it is possible in principle to
obtain a spectrum from a film-covered
sample just by admitting the radiation into
an infrared spectrometer. However, there
are two reasons why this doesn't work
well. First, the radiation is very weak at
room temperature and is emitted away in
all directions. Second, most of the flux is
useless background (black body-like) radiation from the substrate and the film as
well with only a small fractional contribution from the infrared vibrations of interest. Related to the latter point is the fact
that all the parts of the spectrometer itself
(mirrors, apertures, etc.) emit interfering
background radiation. However, strategies have been developed to overcome
these problems, primarily by using samples
at substantially higher temperatures than
the spectrometer and making very accurate corrections for sample background radiation by using film-free reference samples. Commercial instrumentation is available for these experiments for use with
heated samples. Monolayer detection is
not possible by such strategies because of
the large background corrections. However, it is possible through the use of cryogenically cooled spectrometers to detect
monolayers on room temperature samples
(Chabal, 1988; Allara et al., 1984; Chiang
et al. 1987) but such instrumentation is not
available commercially. For cases in which
samples can be heated above room temperature, a major advantage of emission spectroscopy compared to reflection is that the
sample can be of any geometrical shape
and surface roughness. This makes the
method extremely attractive for certain applications such as examining surface films
on odd-shaped machine parts which, in

21.2 Survey of Characterization Techniques

fact, may be normally hot. Of course, such


sample characteristics do not allow quantitative analyses. In general, without extreme effort, emission spectroscopy cannot
be performed in a quantitative manner relative to the cases of transmission and reflection.
Finally, the method of infrared spectroscopic ellipsometry will be mentioned.
This mode is really a special case of the
external reflection experiment in which
changes in the polarization state of the infrared beam are measured rather than the
fraction of beam power reflected, the traditional reflectivity measurement. The major
advantage of the ellipsometry measurement is that the polarization state change
contains information about both the amplitude change and the phase change of the
light whereas reflectivity measurements
only contain intensity information. The
appearance of the additional variable of
the phase shift allows two unknowns of the
sample to be evaluated, for example, sample film thickness and density, whereas reflectivity allows only one, for example, just
sample film thickness with the density
evaluated independently. The calculations
for ellipsometry are based on standard
classical electromagnetic theory. Ellipsometric spectra have the added advantage
that the large amount of data accumulated
from the total numbers of spectral points
can be used to accurately determine wavelength independent sample characteristics
such as thickness, density, surface and interfacial roughness, voids and even density
gradients perpendicular to the surface
plane (depth profiling). This information
is in addition to the normal spectral information involving chemical structure, bond
orientation and other features which affect
vibrational modes and the absorption
cross sections. By utilizing the depth profiling capability, composition gradients

695

could be measured, in principle. The best


demonstrations of the power of ellipsometry as an incisive analysis tool for thin film
structures have been made for spectra obtained in the ultraviolet and visible regions
(see previous section). Here, for example,
thicknesses ideally can be measured to
within 0.1 A or better. Recently, interest in
extensions into the infrared region has accelerated (Roessler, 1987; Ferrieu and Duarte, 1990; Drevillon, 1988) and commercial instrumentation is now available.
However, although it is possible to obtain
spectra of monolayer quantity films, because of signal-to-noise ratio restrictions it
is not yet clear whether these films can be
analyzed quantitatively with present ellipsometry instrumentation. Two studies also
have been reported in which limited polarization measurements have been made on
polymer and Langmuir-Blodgett films
(Graf et al., 1986; Ishida, 1987). For various reasons related to the basic physics of
the experiment, it would appear that IRSellipsometry is most ideal for thin films on
nonmetallic or weakly metallic substrates
as opposed to highly conductive metals. In
addition, since the quantitative value of
the experiment derives from extremely
high precision measurements, only highly
smooth, planar surfaces lead to high precision analyses. However, even for rough surfaces, ellipsometric measurements are of
more value than reflectivity measurements
simply because of the extra information.
Finally, because of the longer wavelength
of the light, the ability of IRS-ellipsometry
to measure thicknesses and perform depth
profiling is significantly less than in the
UV-visible regions, by at least an order of
magnitude. In general, considering all the
above points, it is always an advantage to
use ellipsometric spectra rather than reflectivity spectra if possible, regardless of sample roughness or substrate material type.

696

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

21.2.3 Photons In, Electrons Out


The family of techniques using photons
inelectrons out represents a distinct departure from photon in-photon out techniques with regard to instrumentation and
capabilities. The main instrumental differences arise because electrons are detected.
Accordingly the analysis system must be
under vacuum and variable electric fields
are used for resolving the kinetic energies
of the output electrons when this is required. The main differences in capabilities
center on: surface selectivity, in situ analysis, depth profiling and information content. The generation of the input photon
follows from the types of methods in common use for a variety of applications: Xrays typically from anode emission or synchotrons and UV photons from standard
discharge sources or synchotrons. The
purpose of the photon is to excite an atom
in the target material such that an electron
in some initial energy state is ejected from
the atom into a final vacuum continuum
state where it can be detected. The simplest
process is a direct photoionization of a single electron. More complex processes can
involve multiple electron excitations and
relaxations. A number of these excitations,
both single or multiple, exhibit energies
which are sensitive to the characteristic
structure of the solid, e.g., the valence
states of an element present, and thus can
form the basis of structural analysis methods. Since these input photon beams are
usually not destructive to most common
organic materials except at very high fluxes, the related methods are reasonably
non-destructive. Photons with energies
lower than the UV range are not used since
they are not effective in overcoming the
solid's work function for photoionization
of an electron. In order to obtain spectroscopic information either the vacuum state

kinetic energies of the various electrons


emitted can be analyzed for a constant input photon energy or the input photon energy can be varied and the output electron
current measured without regard to the
electron energies. Both of these methods
can be used to obtain information about
the structure sensitive electronic states in
the analyzed material. The first method is
the basis of XPS, VBXPS and UPS while
the second method is the basis of NEXAFS. The initial interaction mechanism
with the photon always involves a quantum excitation process which creates an
excited electron. However, the secondary
processes which follow on a very short
time scale are extremely important in shaping the capabilities of the analysis. Since
only electrons which escape from the solid
surface can be measured, it is necessary
that the electrons which have been ionized
from their original atoms avoid scattering
events within the solid which would prevent escape. For each material and initial
electron kinetic energy, an average mean
free path will describe the probability of
this escape into a continuum state. For a
photoelectron created at depth z from the
surface of a solid the probability of ejection, /(z), relative to the probability of
ejection at the surface, 7(0) is given by
I(Z)=I(O)[1-Q-ZI[HE)COS0]]

(21-7)

where X (E) is the mean free path of the


electron at kinetic energy E and 6 is the
angle of the electron trajectory from the
surface measured from the normal to the
surface. These values are typically < 10 nm
and hence provide the basis for surface selectivity inherent in all the photon in-electron out techniques. For example, if the
effective total depth probed, de, is defined
as the depth region nearest the surface
which generates 95% of the total possible

21.2 Survey of Characterization Techniques

signal for the entire solid then


dc = 3 X cos 6

(21-8)

These equations can be used to advantage


by varying the takeoff angle 9 for a given
sample and analyzing the intensity response in detail for concentration gradients of the atoms of interest. Finally, it
should be realized that the above mechanism of photon in-electron out implies a
growing charge imbalance which sometimes causes complications in analysis of
polymers since they are usually insulators
and cannot conduct charge away at an appreciable rate. Such charge imbalances alter the electrical potentials in the analysis
system and thus the measured kinetic energies of the photoelectrons can be erroneous.
21.2.3.1 X-Ray Photoelectron
Spectroscopy (XPS)
This technique is designed to measure
the core-level binding energies of constituent atoms in a solid. Perhaps among all the
techniques used in the characterization of
organic surfaces, this is the most popular
one. The popularity arises from several
distinct advantages over other surface
techniques. First, the sample preparation
for XPS measurements is relatively simple,
since solids in different forms (powder,
film or fibers) can be analyzed and the data
interpretation is relatively straightforward. Second, both chemical and elemental information can be provided. Finally,
the quantification of surface species is relatively easy in comparison with many other
surface analysis techniques. Several books
and reviews papers have been written on
the principles and applications of XPS in
materials surface analysis (Siegbahn, 1986,
1990; Nefedov, 1988; Gardella, 1988;
Briggs and Seah, 1990; Powell and Seah,

697

1990; Grunthaner, 1987). By far the majority of applications to organic materials has
been with polymers. However, excellent
analyses have been carried out on supported molecular monolayers and almost any
organic material which is nonvolatile and
can survive moderate X-ray doses would
seem amenable to XPS analysis.
The XPS experiment depends upon Xray induced photoionization of core-level
electrons. The surface to be analyzed is
irradiated with photons from a soft X-ray
source, most commonly Al Ka with an energy of 1486.6 eV but also Mg Ka with an
energy of 1253.6 eV. These two sources are
preferred because of their intrinsically narrow X-ray line widths and the excellent
heat conductivity properties of the two
materials as anodes for X-ray generation
(Briggs and Seah, 1990). A photon of energy hv interacts with an atomic core level
electron of binding energy Eh. If h v > Eh,
a photoelectron can be ejected from the
surface with a kinetic energy Ek based on
the Einstein equation:
hv = Ek + Eh

(21-9)

The process is shown schematically in


Fig. 21-8. Since no two atoms in the Periodic Table have the same exact set of values of their atomic binding energies, XPS
can be used for the purpose of elemental
identification.
Levels
2p (L3)
2p(L 2 )
2s

ls(K)

xps

Figure 21-8. Schematic representation of the photoionization process in X-ray photoelectron spectroscopy (XPS). C Is line is used as the example.

698

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

It was realized quite early in the historical development of X-ray photoemission


that core level electron binding energies
are influenced by the valence state of the
atoms because of the electrostatic interaction between the valence or bonding electrons and the core level electron which is
photoionized (Siegbahn et al., 1969; Hedman et al., 1970). Such interactions result
in a dependence of the binding energy of
the core level electron upon the chemical
state of the atom. Thus the variation in
binding energy of the core level electron is
referred to as a "chemical shift," the magnitude of which may range from a few
tenths to a few tens of eV. This is the basis
of the historical term "electron spectroscopy for chemical analysis" (ESCA). Extensive efforts have been made to measure the
chemical shifts of different functional
groups in organic materials, mostly polymers, and to correlate other properties of
the functional groups with the chemical
shifts (Dilks, 1981; Lopez etal., 1991;
Brito etal., 1991; Beamson etal., 1991;
Gelius etal., 1970; Lindberg etal., 1970).
A data base has been built recently that
includes more than 13 000 individual data
entries up to 1985 (NIST, 1989) and the
new edition of this database is in progress.
Another database of both AES (Auger
electron spectroscopy) and XPS spectra is
being planned by the American Vacuum
Society (AVS) (Powell, 1991). To illustrate
the capability of XPS to identify the functional groups in the surface region of a
polymeric material, the spectrum of a
polyimide thin film is shown in Fig. 21-9
(Atanasoka etal., 1987). The different
functional groups in polyimide can be easily identified from both C Is and O Is spectra. As is quite commonly done, curve resolution has been applied to help resolve
overlapping peaks from two or more component species.

"b
PMDA-ODA

C 1s

2.3.4-0

3
.

292
CO

288

0 1s

284

280

n -> re*

o
*co
CO

'E
o

534

538

530

N 1s

404

400

396

392

Binding Energy (eV)


Figure 21-9. C Is, O Is and N Is photoemission
spectra of clean PMDA-ODA polyimide. The five inequivalent sites in the monomer unit are labeled.
Reprinted with permission from Atanasoka et al.
(1987). Copyright American Institute of Physics.

Even with the large body of literature on


chemical shifts of functional groups on
polymer surfaces gathered during the last
twenty years, it is sometimes still difficult
to make unambigous assignments because
of the small differences in chemical shifts
relative to the inherent instrument resolution. Typical values of the latter can range
from a few tenths to 1-2 eV, but for the
majority of existing instruments, the resolution is much worse than 1 eV. Since shifts
of the most common C l s and O l s core
levels in polymers are within 1 eV, individual peaks from different functional groups

21.2 Survey of Characterization Techniques

often overlap and the whole spectrum suffers from broadening. In addition, the C Is
line from polymers is rather broad by nature (>0.8 eV) (Siegbahn, 1990; Beamson
etal., 1991). To circumvent this problem
one, of course, can use instruments of increased resolution but availability and cost
quickly become significant factors. As an
alternative, chemical labeling or derivatization are quite useful techniques which
have been used by several researchers to
overcome these difficulties. By making use
of specific derivatizing agents which will
react selectively with one functional group
and label it with a distinctive element
(Briggs and Seah, 1990; Chilkoti and Ratner, 1991) an XPS measurement can then
be made of the new element to identify the
presence of the original functional group.
The sensitivity of this method is enhanced
when a new element with a high photoionization cross section is used as the derivatizing agent. The application of chemical
derivatization to the analysis of functional
groups in polymer surfaces has been reviewed recently by Batich (1988). The
reader is referred to this review for detailed
derivatization reactions that can be used to
identify different functional groups and
the original references listed in this review
can be a further source for those who are
considering using this technique to solve
their particular problems.
The simple photoionization process of
core level excitation occasionally becomes
complicated as other electronic level
changes can occur simultaneously with the
primary excitation. In particular, for unsaturated compounds, higher lying restates can be simultaneously excited to n*
states at the expense of lowering the photoelectron kinetic energy. This results in a
so-called 71->TC* shake-up peak (Gardella
etal., 1986). The latter appears at about
10 eV higher binding energy than the pri-

699

mary core level peak. This unique feature


can be utilized for the identification of the
structure of aromatic polymers and sometimes a quantitative determination of the
amount of a particular aromatic containing function or polymer block unit in the
polymer surface can be made (Lopez et al.,
1986; Gardella etal., 1984; Clark and
Dilks, 1977). It has been reported that the
n-*n* transition can be used to follow the
destructive effect of X-ray or ion bombardment on polymers containing aromatic rings (Storp and Holm, 1979; Marietta
et al., 1991; Briggs and Hearn, 1985). The
disappearance of the 7i->7t* satellite after
X-ray or ion bombardments is a convenient signal of the destruction of the aromatic structure of the polymers.
X-ray excited Auger lines (see Sec.
21.2.4.1) can provide additional information for the identification of chemical functional groups or chemical states (Reilley
and Everhart, 1982). Auger signals often
show binding energy shifts significantly
large than their XPS countparts. X-ray excited Auger lines have been used to identify the chemical states of some metallic materials that otherwise can not be distinguished by core-level lines. A typical example of such an application is the identification of the different oxidation states of Cu.
There is virtually no difference in the Cu
2p binding energies of metallic (Cu) and
cuprous oxide (Cu + ) whereas a large
chemical shift between cuprous oxide and
metallic copper of about 2 eV is observed
in the Auger spectrum (Strohmeier et al.,
1985). However, Auger lines have been
used mainly in inorganic materials although Wagner etal. (1980) have shown
that Auger spectra can be used in the study
of oxygen-containing polymers.
One of the main advantages of XPS over
other surface analysis techniques such as
AES, SIMS and ISS is the relative ease of

700

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

quantitative measurements of surface


composition. To some extent this is inherent in the simple physical mechanism of
the method but to a great extent is the
result of extensive efforts during the last
twenty years to provide the necessary calibration data for quantification. Several review papers have been written to deal
specifically with this aspect (Powell and
Seah, 1990; Ebel et al., 1988; Seah, 1986,
1991; Battistoni et al., 1986; Grant, 1989)
and the interested reader is referred to
those reviews for significant levels of details.
There are basically two approaches for
quantitative analysis in XPS: (1) the elemental sensitivity factor method and (2)
the first principles method. In the elemental sensitivity factor approach, the photoelectron intensity of a strong line of a
desired element is determined relative to
another standard element (in most cases,
fluorine) with both elements measured in
the form of pure standard compounds
(Wagner, 1977; Berthou and Jorgensen,
1975). Such sensitivity factors are then
used for quantitative analysis using the following equation:

ment and the element whose sensitivity


factor is to be measured. Further, the sample should be atomically homogeneous
and ideally should be fractured in ultrahigh vacuum just prior to analysis in order
to provide a pristine surface (Powell and
Seah, 1990). Unfortunately these conditions are usually not practical. In most sensitivity measurements, inorganic powders
with the highest commercially available
purity are used along with polymers of exactly known compositions (Strohmeier,
1991). The problems associated with surface contamination, sample homogeneity
and radiation damage must be addressed
in measuring the elemental sensitivity factors in a particular instrument. Several sets
of elemental sensitivity factors have been
published in the literature (Briggs and
Seah, 1990; Berthou and Jorgensen, 1975;
Wagner, 1972; Ward and Wood, 1992).
In the first principle approach, the measured photoelectron current from a core
level of atom A for a homogeneous surface
can be related to the atomic concentration
of A by the following equation (Grant,
1989),
(21-11)

(21-10)
X () cos 9 F (, A) T(E, EA) D (, A)
where Ct is the atomic fraction of the element i, / is the measured signal intensity
and 7 is the sensitivity factor.
Because of the instrument dependence
on these measurements, the elemental sensitivity factor approach is only accurate if
the sensitivity factor is obtained on the
same instrument (or the same type of instrument) under exactly the same conditions as those for the actual unknown determinations. However, this often presents
a problem since the reference samples
should contain at least one standard ele-

where:
7A = photoelectron current at the detector;
$ = X-ray flux at a characteristic energy;
NA = number of atoms of element A per
unit volume;
at = total photoionization cross section;
P = asymmetry parameter;
a = angle between the input photon direction and the ejected photoelectron
direction;
8 = angle between the photoelectron direction and the surface normal;

21.2 Survey of Characterization Techniques

X (E) = total inelastic mean free path at the


kinetic energy of the photoelectrons;
F(E, EA) = electron-optical factor relating
the photoelectron current emitted at
kinetic energy E to the current of the
electron entering the analyzer at kinetic energy EA;
T(E,EA) = analyzer transmission function;
D (E, EA) = efficiency of the detector.
Because the above equation involves the
knowledge of the geometry of the analyzer,
the detector, the signal analysis electronics
and the processing of the spectra (Seah,
1980), it is not possible to do quantitative
analysis without a quantitative characterization of the instrument. The latter is usually not available which provides a severe
difficulty with the first principle approach.
For this reason sensitivity factors are preferred for doing quantitative XPS.
The quantitative detection limit, or surface sensitivity threshold, in XPS is in the
range oflO~ 2 -10~ 3 atomic fraction content in the solid (Hoffmann, 1986) for photoelectrons with kinetic energies typically
produced using AL Ka or Mg Ka sources.
For these cases, one can calculate that the
total signals come from depths of < 10 nm
using Eq. (21-8). The surface selectivity
can be further enhanced by varying the
take-off angle with respect to the sample
surface normal, as shown in Fig. 21-10,

701

and the corresponding effective total depth


probed can be calculated from Eq. (21-8) if
X is known.
It is clear from the above analysis that
correct values of X are essential for quantitative work and significant effort has been
expended for a number of years to obtain
reliable values for different materials and
photoelectron kinetic energies (Atanasoska etal., 1992; Bain and Whitesides,
1989; Cartier etal., 1987; Clark etal.,
1981; Laibinis etal., 1991). However, the
data reported exhibit wide scattering
which is believed to be due to difficulties in
preparing uniform and homogeneous organic films and in determining accurately
the thickness of the formed films. It has
been shown recently that self-assembled
monolayers (SAMs) are extremely homogeneous organic films with thicknesses
which can be controlled on the angstrom
scale (Nuzzo etal., 1990; Porter etal.,
1987). Using these films, Laibinis etal.
(1991) have proposed the following expressions for the dependence of X (nm units) on
the kinetic energy of photoelectrons in organic films:
X = 0.031 EA65

(21-12)

or
X = 0.9 + 0.0022 EA

(21-13)

Based on above equations, the X value of


C1 s photoelectrons at an energy of

Analyzer
X-ray Photons

d=

X-ray Photons

I
d' = 1 cos

Figure 21-10. Schematic representation of the effect of take-off angle on the depth probed in X-ray
photoelectron spectroscopic (XPS)
measurements.

702

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

1201.6 eV (an Al Ka source) in a dense


organic medium, similar to a self-assembled hydrocarbon monolayer, is about
3.5 nm. Using this value of X and 0 = 45,
rfe7.4nm; for 0 = 0, d ^ l 0 . 5 n m . In advanced instruments with high X-ray fluxes
and efficient detectors, it is possible to set
the instrument to very low 6 values and
force de down to values of ^0.5 nm.
The change of probing depth by alteration of take-off angle has become the basis of a non-destructive technique to obtain
depth profile information. This application was first suggested by Fadley and is
referred to as angle-resolved X-ray photoelectron spectroscopy (ARXPS) (Fadley,
1984), or angle-dependent X-ray photoelectron spectroscopy (ADXPS) (Dwight
et al., 1990). This approach has been used
by several researchers to obtain compositional depth profiles in polymeric materials (Gardella, 1988; Dwight etal., 1990;
Gerenser, 1988; Bhatia etal., 1988). The
increased surface selectivity is critical in
addressing the problems of preferential
surface enrichment, surface segregation
and surface treatments of polymeric materials. Dwight etal. (1990), for example,
have used ADXPS to study the surface
segregation of the polysiloxane component in graft copolymers of polyesterpolysiloxane, polycarbonate - polysiloxane and poly (ethyl methacrylate)-polydimethylsiloxane. Significant enrichment
of the siloxane component is observed at
the surface of these copolymers.
In principle, the lateral resolution of an
XPS measurement could be in the
nanometer scale, the diffraction limit of
the X-ray beam, but, in fact, because of the
difficulty in practice in developing focused
beam instruments (Kelley, 1991), the typical spot size has remained in the millimeter
scale. However, there has been significant
and growing demand for the capability of

analyzing small areas in many application


such as microelectronics fabrication (Ninomiya etal., 1991). Fortunately, significant progress has been seen over the last
few years and it is now possible to obtain
spectra with a lateral resolution of 5 25 \xm (Beamson et al., 1990; Coxon et al.,
1990; Gelius etal., 1990; Pianetta etal.,
1990; Seah and Smith, 1988). However, in
order to maintain the output electron signal at a level for reliable detection and
good energy resolution, the X-ray fluxes in
the small area focus must be quite high and
only a limited number of organic materials
are able to resist radiation damage during
analysis (Kelley, 1991). Inorganic materials however, are usually much more radiation resistant and thus present less of a
problem in general. Typically, XPS analysis from conventional instruments with lateral resolution no better than 100 Jim is
considered to be nondestructive for polymers. This is generally true in comparison
with other surface analysis techniques,
e.g., AES, ISS or SIMS (excluding time-offlight detection). However, damage does
occur in some sensitive polymeric materials and studies of the radiation damage of
polymers during XPS have been reported
for poly (vinyl chloride) (Akhter et al.,
1988), poly(methyl methacrylate) (Pan
etal., 1991; Buchwalter and Czornyi,
1990) and perfluoropolyether (Barth et al.,
1988). Generally, new polymers should be
analyzed as a function of X-ray intensity
and exposure time to see if degradation
is occurring over the period of analysis.
Newer instruments have much more efficient electron detection so this decreases
analysis times with no loss in signal to
noise thus alleviating some of the degradation problems.
Since most organic materials are insulators, a pervasive problem with their XPS
analysis has been surface charging which

21.2 Survey of Characterization Techniques

shifts the electrical potential of the sample


with respect to the detection system. This,
of course, shifts the energy scale of the
spectrum. However, this shift is easily corrected in practice by various means, most
commonly by using the ever-present carbon contamination as a reference; most
workers set this C l s line at 285.0 eV. A
more serious complication is caused by
nonuniform charging which results in distortion and broadening of the peaks in the
spectrum, making data interpretation difficult or even impossible (Cros, 1992; Barr,
1989; Barth et al., 1988). Some samples are
more susceptible than others to this phenomena and the reasons are not well understood. Surface charging is especially
severe in XPS instruments using monochromatized X-ray sources. The reason for
this is that X-ray anode sources produce a
steady background flux of low energy electrons as the bremsstrahlung X-rays strike
the X-ray gun window and the internal
parts of the apparatus and these electrons
then strike the sample, partially compensating for the depletion of charge by the
photoelectron emission. In a monochromatized instrument, the monochromator
system prevents the background electrons
from the source from reaching the sample.
Taking advantage of this neutralization
principle, electron flood guns have been
designed and placed near the sample to
purposely supply an adjustable flux of low
energy electrons for surface charge compensation (Cazaux and Lehuede, 1992).
Since its development in the late 1960s
and early 1970s, XPS has been used extensively for organic materials, mostly polymers. The biannual review series on XPS
and AES in Analytical Chemistry and other review sources have provided numerous
examples of the application of XPS in the
area of polymer surface and interface analysis (Turner, 1988, 1990, 1992; Gardella,

703

1988; Dilks, 1981). A series of more than


30 papers has been published by Clark
et al. exploring the application of XPS to a
variety of polymeric materials (Clark and
Harrison, 1981). Because of the high surface sensitivity and the ability to provide
information about the surface functionalities, XPS has been used extensively in
the analysis of surface-treated polymers
(Gerenser, 1988; Occhiello etal., 1991;
Lazzaroni etal., 1991; Pawson etal.,
1992). XPS has also been used in the investigation chemical interactions between
vaccum-deposited metals and polymer
surfaces (Bou etal., 1991; Pertsin and
Pashunin, 1991; Atanasoska etal., 1990;
Burkstrand, 1981; Ho et al., 1985; JordanSweet etal., 1988; Mack etal., 1990;
Chtaib et al., 1991; Anderson et al., 1993).
Ho et al. (1985), for example, have used
XPS to study the chemical interaction between aluminum metal and a polyimide
surface by in situ measurement of the Cls,
O Is, N Is and Al 2p lines during thin aluminum layer deposition. The appearance
of low binding energy peaks in C l s and
N Is spectra indicate that both Al-C and
A l - N complexes are formed at the aluminum-polyimide interface. Similar chemical interactions have been observed in
other systems (Goldberg et al., 1988; Jordan-Sweet et al., 1988). XPS has also been
used to study the surface of peeled-off interfaces in order to gain some understanding of the failure process. Such studies are
important to understanding the relationships between interface bonding and mechanical performance in composite materials (Watts, 1988).
It is clear that XPS will continue to be
used extensively in the future for the study
of the surfaces and interfaces of polymeric
materials because of all the advantages discussed above. The evolution of new instruments with higher resolution (state of the

704

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

art ~0.3 eV) and higher X-ray power


(8 kW) (Beamson et al., 1990,1991; Gelius
et al., 1990), which translate into increased
capabilities of distinguishing different
functional groups and shorter acquisition
times, will further enhance XPS for the use
in organic materials.
21.2.3.2 Valence Band X-Ray
Photoelectron Spectroscopy (VBXPS)
and Ultraviolet Photoelectron Spectroscopy
(UPS)
The chemical dependence of core-level
binding energies has been demonstrated to
be of great utility in the chemical characterization of surfaces in the previous section. However, there are definite limits on
XPS and further details must be obtained
in other ways. In this sense it is a distinct
advantage to gain access to information
about valence bands of a solid since these
are intimately involved in the formation of
chemical bonds and thus are far more sensitive to subtle details of chemical structure
than are the core levels. The two traditional means of generating valence band information at surfaces utilize either X-ray photons (VBXPS) or UV photons (UPS) to
generate transitions between the various
VB levels. Although these two techniques
are closely related they are presented separately below in the order of VBXPS and
UPS. Most of the principles are presented
in the VBXPS discussion and the additional considerations for UPS are presented at
the end.
Experimentally, valence band XPS is
quite similar to XPS. The only essential
difference is that the photoionization process in the former involves X-ray excitation of an electron from a valence band
initial state rather than a core level state.
The only actual difference is that in the VB
measurement, the kinetic energy of the

ejected photoelectron is higher than for a


core-level excitation because for the higher
lying valence band state a much smaller
fraction of the input X-ray photon energy
is required to overcome the work function
of the surface than would be required for a
deep lying core level. Most XPS spectrometers can easily accommodate analysis of
these higher electron energies so VBXPS
spectra may be obtained from the same
instrument as XPS spectra. However, on a
finer scale there are important instrumental differences. Because of the high density
of states in typical valence bands, the spectra have much more fine structure than
core level spectra. In order to optimally
resolve the fine structure, monochromatized or synchotron X-ray sources are
needed. The major problem with unmonochromatized radiation is that the
K al 2 excitation line has a low intensity
satellite with an offset of 10-12 eV, which
interferes with the spectra in the valence
electron region from 0 to 35 eV (Pireaux
et al., 1981). In addition to the fine structure, VB photoelectrons also have intrinsically low cross sections for ionization and
thus are produced at weak signal levels
(Boulanger et al., 1989). For this reason it
is important to reduce the bremsstrahlung
continuum of the radiation source in order
to avoid background interference from the
secondary electrons.
The major advantage of VBXPS is the
sensitivity to the chemical structure of the
surface region. This is easily understood by
the fact that, in contrast to XPS in which
the emitted core level electrons are detected, in VBXPS the valence electrons that
are directly involved in chemical bonding
are detected. Thus VBXPS has a great potential as a tool for probing both the electronic valence band structure and for
studying chemical bonding and reactions
at surfaces. The schematic of both XPS

21.2 Survey of Characterization Techniques

705

Binding energy of electrons


X

"in

o
v

levi

U)
01

0)

'

>

>

(N *~ *$
>

'N

(D

^-^
D

T3

X)

d)

(D
a

"a

o
o
o
c

c
>

(D

ipi

\\i

u ; ;

_J

*-

^^ *
J
'N ^
\
\

Intramolecular
electronic
relaxation
Intermolecular
electronic
relaxation

*- i o

A V\
v n '1

> | r-

Figure 21-11. An idealized


photoelectron spectrum relative to a hypothetical energy level scheme. The oneto-one correspondence
between spectral peak and
the energy levels in the relevant feature is shown.
Reprinted with permission
from Salaneck (1985).
Copyright CRC Press, Inc.

Kinetic energy measured


(wide range XPS spectrum)

and valence band spectra and the relation


with the electronic structure of material is
shown in Fig. 21-11 (after Salaneck, 1985).
Of specific interest is the fact that for
organic materials many of the differences
between samples and surface treatments
are often quite subtle in terms of structures
involving such effects as hydrogen bonding, isomerization, group orientation,
crosslinking or conformational sequences.
Most of these secondary types of effects
generally cannot be discerned by core level
spectroscopy but may be by VB spectroscopy. The interpretation of both primary
and secondary effects in polymers has been
studied to a limited extent (Pireaux et al.,
1981; Orti et al., 1990). The application of
VBXPS in polymeric materials have been
reviewed recently by Salaneck (1991) and
Pireaux et al. (1990) and the reader is referred to these publications for a more detailed discussion.
In spite of the strong potential of
VBXPS as a highly sensitive probe of
chemical structure, the problems discussed
above have prevented its extensive use. In
order to overcome the low VB photoionization cross sections, very long acquisition times are required for an acceptable
signal-to-noise ratio in a traditional XPS

instrument (Pireaux et al., 1981). Newer


instruments with improved electron analyzers, multichannel detectors and higher
X-ray fluxes from rotating anodes or synchotrons have improved this situation considerably. However, the problem still remains that exposure of an organic sample
to intense X-ray radiation for prolonged
period of time implies a higher probability
of radiation damage, as discussed above
for XPS. Thus extreme care must be taken
to identify whether any effects of radiation
are present. A second major problem arises from the complex nature of the valence
band spectra. It is, of course, this complexity which provides, in principle, rich chemical detail about the surface, but as a consequence the interpretation of spectra is
not straightforward. In general, it is necessary to perform molecular orbital calculations of the valence bands in order to
provide a basis for the interpretation of the
complex spectra (Boulanger etal., 1989,
1991; Orti etal., 1990). Fortunately, with
the continual improvement in computing
capabilities and electronic structure algorithms, such calculations are becoming
more straightforward, although not yet at
the level to entice the standard user to do
them.

706

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

Even with the difficulties as discussed


above, the unique fingerprinting nature of
VBXPS has been applied by several groups
to investigate a variety problems in polymers. The uniqueness of this technique can
be demonstrated by the observation that
different polymers which have the same
core level spectra show totally different valence band spectra. Figure 21-12 illustrates
this point with the three similar hydrocarbon polymers polyethylene, polypropylene
and poly(l-butene) (Briggs, 1989). Several
sets of polymers which also illustrate this
point include: polystyrene-poly(/?-phenyl),

polypropylene oxide - poly (vinyl methyl


ether), poly(>z-butylacrylate) - poly(7-butyl acrylate)-poly(7-butyl acrylate) and
poly (cis -1,4 - butadiene) - poly (1 - butene)
(Pireaux etal., 1981, 1990; Castner and
Ratner, 1990; Stickle and Moulder, 1991).
It has been observed that VBXPS not only
distinguishes polymers with different compositions but also polymers with the same
chemical formula but different bond organizations, i.e., constituent isomers. Several
examples can be found in the literature
showing the unique capability of VBXPS
to distinguish fine structure in polymers
such as head-to-head, head-to-tail linkages, stereoisomers and tacticities (Pireaux
et al., 1981). The XPS valence band spectra of isotactic and syndiotactic polypropylenes are shown in Fig. 21-13. However, as might be expected, the unique "fingerprinting" feature of VBXPS is always
accompanied by the difficulty in a fundamental interpretation of the spectra. Theo-

/
/

Isotactic
Syndiotactic

25
Binding energy (eV)

Figure 21-12. Valence band spectra from the hydrocarbon polymers: low density polyethylene (LDPE),
polypropylene (PP) and poly(butylene) (PB). All these
polymers give identical C Is spectra. Reprinted with
permission from Briggs (1983), in Practical Surface
Analysis (Eds.: D. Briggs, M.O. Seah). Copyright
John Wiley & Sons, Ltd.

40

J j

30
20
10
Binding energy (eV)

Figure 21-13. Valence band spectra of iso- and syndio-polypropylenes. Reprinted with permission from
Pireaux etal. (1981). Copyright American Chemical
Society.

707

21.2 Survey of Characterization Techniques

retical simulations have been performed


on a variety of polymers to assist the interpretation of spectra (Boulanger et al.,
1989, 1991; Orti et al., 1990; Pireaux et al.,
1990; Delhalle et al., 1987; Cain, 1988).
VBXPS has also been used quite extensively in the study of the electrical conductivity
of polymeric materials in an effort to correlate the electronic structure of the polymers with the electrical conductivity
(Pireaux et al., 1990; Monkman et al.,
1989; Obrzut et al., 1989). The application
of VBXPS in the study of conducting polymers has been reviewed by Salaneck
(1986). For applications of VBXPS outside
the general area of organic materials, such
as superconductivity, corrosion etc., the
reader is referred to the work by Sherwood
and his colleagues (Sherwood, 1991;
Thomas etal., 1990).
Valence bands can also be probed by the
use of UV photons to cause ionization of
the valence electrons. Essentially the same
information is obtained as with the soft
X-ray probes of VBXPS but major differences occur in the cross sections of photoionization of the different valence states
so that UPS spectra often bring out transitions which are hard to observe with
VBXPS, while, of course, the opposite may
occur for other transitions. Thus, it is an
advantage in obtaining VB spectra to be
able to use both types of photon probes
and they both can be installed on the same
instrument easily. UV photons with energies typically less than about 50 eV are
used. Line sources (atomic emission) have
an advantage in that they intrinsically have
very narrow line widths on the order of
about a few meV FWHM (Pireaux et al.,
1981; Salaneck, 1981). Accordingly, detailed valence band spectra can be expected. However, the potential of higher resolution in UPS is not realized in studying
solid state organic materials, primarily

polymers, since the molecular orbitals are


grouped into bands, broadened by vibrational excitations such that the line widths
of a valence band peak is w 1 eV near
room temperature (Salaneck, 1981; Duke
etal., 1978), a value considerably larger
that the excitation line width. A comparison of the valence band spectra of
polystyrene from a monochromatized Al
Ka source and a monochromatized He II
line source at 40.8 eV is made in Fig. 21-14
(from Salaneck, 1985). UPS, however,
does offer one distinct advantage over
VBXPS. The low kinetic energy (0-40 eV)
of UV-excited photoelectrons results in a
small mean free path of the photoelectron
and thus the escape depth is very
small, ~ 0.5 nm. UPS is therefore much
more surface selective than XPS or
VBXPS. However, a consequent problem
which arises is that surface contamination
easily can dominate the observed spectra
with such small analysis depths so ultraclean samples and analysis chambers are
necessary for accurate work. This can be a
particular problem with polymers and

Solid phase: polystyrene


XPS:
monochromatized
AI(KJ radiation

Pireaux, et al.

UPS:
monochromatized
He I I radiation
40 8 eV

Salaneck, et al.
20 15 10 5
Binding energy (eV)

Figure 21-14. Comparison of valence band spectra


obtained from monochromatized Al Ka and He II
radiations. Reprinted with permission from Salaneck
(1985). Copyright CRC Press, Inc.

708

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

molecular films since these surfaces cannot


be cleaned by traditional surface methods
such as sputtering without extensive damage to molecular structure.
21.2.3.3 Near Edge X-Ray Absorption Fine
Structure (NEXAFS)
In the previous section it was shown that
with the techniques of UPS and VBXPS,
UV and X-ray photons can be used to excite transitions within the valence band
and thus generate spectra which contain
valuable information about chemical and
electronic structure in organic materials. It
is also possible to obtain information
about valence band levels by using X-rays
to excite electrons from deep core levels
into the valence band. A synchotron-based
technique which utilizes this approach is
NEXAFS or XANES. This technique can
be considered as an alternate method of
obtaining valence band information and
has been useful for the characterization of
the electronic structure and orientation of
functional groups in the surfaces of polymers and thin molecular films on solid supports. The mechanism of the spectroscopy
involves absorption of an X-ray with excitation of a K-shell electron to unoccupied
orbitals just below the ionization
threshold, thus the term near edge X-ray
absorption fine structure. When the excited state relaxes, X-ray radiation can occur
but also a secondary electron can be ejected. Since the electron emission is localized
near the surface as determined by its mean
free path, it can be detected to provide a
surface selective measurement, as with other electron-out spectroscopies. Without regard to the specific energy of the secondary
electron emitted, the simple observation of
the emission can be taken as a signal that
a K-shell electron was excited to an unoccupied state in the initial X-ray absorption.

However, in order to investigate the energy


distribution of the unoccupied valence
band states this way it is necessary to
change or tune the input X-ray energy. In
practice this is done with a synchotron
source. Thus NEXAFS spectra of surfaces
are obtained by measuring the secondary
electron yield as a function of the incident
photon energy. It has been shown that
NEXAFS can be useful in the study of
the orientation and bonding of small
chemisorbed molecules on solid surfaces
and in the investigation of LangmuirBlodgett films (Solomon etal., 1990;
Skotheim etal., 1989; Rabe etal., 1988;
Outka et al., 1988; Friend, 1989). For applications to the study of molecular structure and orientation of a variety of
molecules on the surface of solid prior to
1987, the reader can refer to the review
article by Stohr and Outka (1987). Unlike
XPS which has found extensive applications to polymer surface analysis, NEXAFS has seen only limited applications in
this area. The reason for this is due primarily to the fact that a tunable X-ray source
is required and only a synchotron has this
capability on a routine basis. In addition,
as expected for a VB spectroscopy, the
spectra are often quite complex. Polymers
which have been studied include polystyrene (Outka and Stohr, 1988),
polyethylene (Rabe et al., 1988) and polyimide (Jordan-Sweet etal., 1988; JordanSweet, 1990; Kowalczyk and JordanSweet, 1990). NEXAFS also has been used
to study chemical reactions between polymers and metals (Jordan-Sweet et al.,
1988; Jordan-Sweet, 1990; Kowalczyk and
Jordan-Sweet, 1990). The systems studied
include aluminum deposition on polyimide
and the adsorption of polyimide on a variety of solid metal substrates. These studies
suggest that NEXAFS can be a powerful
tool for studying bonding interactions and

21.2 Survey of Characterization Techniques

molecular orientation of fairly complicated materials systems (Jordan-Sweet, 1990).


However, as with VBXPS and UPS, accompanying molecular orbital calculations
need to be done in order to obtain detailed
spectral assignments and quantification of
experimental data. Further, X-ray damage
is always a potential problem for many
organic materials.
21.2.4 Electrons In, Electrons Out

With a change from a photon as the


input probe particle to an electron, the
fundamental nature of the probe mechanisms involving excitation and scattering
changes dramatically. The momentum,
charge and fundamental quantum statistical properties of the two probes differ significantly and change both the character of
the analysis experiments and the structural
information which can be obtained from
them. There are two basic types of experiments: externally generated electron
beams with the source removed at a
macroscopic distance from the target and
near-surface generated electrons with the
source only atomic distances away. The
first category of methods includes AES,
EELS, HREELS, LEED and TED. The
second category includes the tunneling
methods, IETS and STM. In the overview
below each category is discussed in order.
Because of the close relationship between
STM and the technique of atomic force
microscopy (AFM), a technique which is
not based on an electron probe, the two
techniques are discussed together in the
STM section. Separating the discussions
would keep the organization of this chapter rigorously coherent but, on the other
hand, would diminish the usefulness of the
presentation since the two techniques are
so highly complementary, both in terms of
instrumentation and imaging capabilities.

709

From the point of view of instrumentation in the AES, EELS, HREELS, LEED
and TED experiments, the input portion
requires potential field gradients to accelerate electrons to the desired kinetic energies and electron optics to create a beam
with the desired spatial organization and
direction. In some sense, this is advantageous since a wide range of electron energies can be controlled in a similar way, in
contrast to manipulation of photon energies with varying sources, optical materials
and physical mechanisms. Of course, it follows from the fact that electron beams
must be freely propagated that a vacuum
environment is essential and thus in situ
capabilities for liquids or gases are not
possible as with photon-based techniques,
so in this sense the use of electron beams
is a disadvantage compared to photons.
From the point of view of the initial collision process, the momentum and energy
exchange processes with the target atoms
are very different than for photons. First
of all, the collision mechanisms are more
complicated but as a consequence can offer more information; in fact, in principle
the electron in-electron out techniques
can be considerably more informative than
the photon in-photon out techniques. The
input electrons can induce a multiple series
of internal excitations and decays of which
the most important one is the Auger process, the basis of AES. They also can undergo a one-step energy loss by excitation
of electronic or vibrational states which
gives rise to the electron energy loss or
EELS spectroscopies. Since the spacings
between vibrational states are relatively
small (an order of magnitude or so) compared to electronic spacings, high resolution measurements must be used to obtain
vibrational spectra with these methods and
only EELS can provide this resolution
which in this high resolution mode leads to

710

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

the term high resolution electron energy


loss or HREELS. Further, as electrons are
accelerated to higher kinetic energies their
wavelengths become shorter and can be
easily controlled to be of the approximate
magnitude of atomic spacings so that diffraction and imaging of materials can be
carried out. These effects give rise to the
LEED and TED diffraction experiments
and the very common imaging technique
of transmission electron microscopy
(TEM). Secondly, for electron and photon
beams of equal energies directed at the surface of a bulk solid, the electrons will penetrate much less deeply into the solid. This
is because of the fact that upon penetration, the electron, relative to the photon,
will scatter much more strongly off the existing electrons in the solid. These scattering processes lead to a large number of
secondary and further events in the solid
matrix as determined by the initial electron
energy. This creates a problem since, for
typical electron energies, the secondary
events are mostly electron ionization from
bonding states which usually cause irreversible, destructive chemical changes to
organic materials. These considerations
are important for the external electron
beam experiments, AES, EELS, HREELS,
LEED and TED. Therefore, unless extremely low currents are used in these
methods, the analyses usually will not be
useful for detailed information about organic surfaces.
For the remote detection techniques,
AES, EELS, HREELS, LEED and TED,
the detection mechanisms involve the ejection of an electron from the sample into
the vacuum and thus are subject to the
considerations put forward in previous
sections. In particular, since the kinetic energies of the ejected or scattered electron
are in the range discussed for the photon
inelectron out techniques, it is clear that

the escape depths will be the same. Thus


the electron inelectron out techniques are
all quite surface selective, and this, of
course, is an advantage in surface analysis.
Another important aspect is that the ability to focus electrons is very developed, for
example, TEM involves atomic resolution
beams, and accordingly the imaging capabilities of electron probe techniques are
quite good, in the nanometer scale provided signal-to-noise ratio is achievable. A
summary of the above information is that
the electron in-electron out processes with
remote detection are: easy to control, surface selective and highly sensitive to structural detail but not always likely to provide
useful analyses because of complexities in
the mechanisms and problems with materials degradation.
The IETS and STM experiments, in
contrast to the above ones, are completely
different with regard to the instrumentation and the physical mechanisms. Both of
these techniques place a voltage-biased
electrode within angstroms of a conducting substrate in order to allow electrons to
tunnel between the substrate and electrode
through an intervening ultrathin film, surface layer of interest or an open gap. In the
case of IETS the source and sample layer
are integral parts of a small, solid state
device which is fabricated for each measurement. In both techniques the energies
of the tunneling electrons are controlled by
biasing the source, a planar electrode or a
tip, at low voltages with respect to a
ground plane which supports (or can be)
the sample. Since the interaction mechanisms with the molecules of interest involve passage of a tunneling electron to (or
from) the ground plane, it is imperative
that these molecules be in a layer which
allows the tunneling, usually accomplished
by making the layer of 1 nm or less
thick. The interaction mechanisms may be

21.2 Survey of Characterization Techniques

as simple as just allowing tunneling in order to see if molecules are there or not at
each point on the surface (imaging by
STM) or as complex as coupling with vibrational states to obtain vibrational spectra (IETS).
21.2.4.1 Auger Electron Spectroscopy
(AES)
Auger electron spectroscopy has been
one of the well-established techniques for
surface analysis over the years. Of the
three traditional surface selective techniques, AES, XPS and SIMS, it has been
stated that AES accounted for 50 % of all
three types of surface instruments worldwide in the early 1980s (Seah, 1984). In
spite of this popularity, AES has not found
many applications in organic materials,
primarily because of sample charging and
electron beam induced degradation. However, AES does have unique capabilities
and deserves mention in this review.
The basic principles have been discussed
in several recent review papers (Ramaker,
1991; Chang, 1987; Paterson, 1987; Seah
and Briggs, 1990). The mechanism is
rather unique and depends upon multiple
inner-shell transitions. The sample is bombarded by an electron beam of 1-10 keV
energy in order to eject core electrons from
a level Ex to another state of arbitrary energy. The core hole is then filled by an
internal process in which an electron from
a level of energy Ey falls into the core hole.
In order to balance the energy, a third electron of energy Ez is ejected simultaneously
from the atom. This last electron, called an
Auger electron, departs into the continuum with a kinetic energy Ea given approximately by
p

_ p

. p _ p

(21-14^

The important aspect of this equation is


that the Auger electron energy is indepen-

711

dent of the energy of the input electron (or


photon in X-ray induced Auger emission)
as long as the energy is sufficient to produce a core hole. Thus the Auger electron
energy is directly characteristic of the
target atom. These energies range from
0-2 keV and are highly characteristic of
the specific element. For this reason Auger
spectroscopy is extremely useful for elemental identification, similar to XPS. Further, because the mechanism involves escape of an electron from the surface into
the vacuum, there is a characteristic escape
depth associated with the process and thus
the spectroscopy is surface selective. For
the range of kinetic energies of 0-2 keV,
the effective analysis depth is <10nm
(Seah and Briggs, 1990). One significant
advantage of using electrons as a probing
beam is that it is relatively easy to focus an
electron beam to a very small size. AES
thus has a great advantage over XPS in
small spot analysis, since the lateral resolution of an AES instrument can be as good
as 3-5 nm (Powell and Seah 1991; Takutaka et al., 1987). This makes it a routine
to do surface analytical imaging, a technique often referred to as scanning Auger
microscopy, SAM (Frank, 1991; Browning, 1987; Kelley, 1991). Auger electrons
come from inner shell levels, and as with
XPS, these levels are sensitive to chemical
bonding to a measurable extent. Thus,
AES can provide chemical bonding information, which can be obtained via measurements of peak shifts and curve shapes
(Ramaker, 1991; Paterson, 1987). However, the information content is diminished
because of damage to the surface caused
by the secondary electron events (Chang,
1987). Such damage is especially severe for
organic materials and has been a limiting
factor in the application of this technique
in the analysis of organic materials as testified by the lack of examples of applications

712

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

in the literature. Some polymers have been


studied and these include polyvinyl chloride (Cota et al., 1986) and styrene butadiene rubber (Lin, 1987). The majority of
applications of AES to organic materials
have been to the depth profile analysis of
metal-polymer interfaces using the combination of AES with ion sputtering (see
Sec. 21.2.7). Examples include the interface analysis in TbFeCo-photopolymers
(Hashimoto, 1991), Cu-Ti-polyimide
(Furman et al., 1990) and Al-polypropylene (Andre et al., 1990, 1989) structures.
AES-ion sputtering offers some advantage over sputter -XPS depth profile analysis because typically at least a hundredfold or so smaller area can be analyzed by
AES than by XPS. Primarily for this reason, depth profiling is now the most important mode of Auger analysis. Finally it
must be mentioned that the heavy flux of
electrons entering and leaving the sample
is never equal because of secondary processes and so the excess charge must be
immediately leaked off or variations of the
electrical potential of the sample will cause
serious errors in the energies of the Auger
lines. Since most organics are insulators,
the charging problems become severe and
are difficult to compensate (Chang, 1987).
21.2.4.2 Electron Energy Loss
Spectroscopy (EELS)
In an electron energy loss experiment,
an electron of fixed energy (monochromatic so to speak) scatters inelastically from
an atom or group of atoms with the loss in
kinetic energy transferred exactly into an
excitation of a quantum state associated
with the scattering centers. In EELS, an
acronym traditionally reserved for electronic excitation, the sample is irradiated
with a high energy electron beam, usually
100 keV or higher. In many cases irradia-

tion is done in the transmission mode using thin film samples; otherwise it is done
in reflection from the surface (Strydom
and Hofmann, 1991). The impinging electrons cause a large number of different
processes to occur and a variety of energies
arise in the output flux. However, for
EELS only a selected momentum and energy range of the scattered electrons is
measured in order to concentrate on the
energy losses resulting from inner shell
ionizations, inter- and intraband transitions and plasmon excitations. The electron energy losses from these processes
range typically from 1 to 24 eV. These energy losses are characteristic of the specific
atoms present in the specimen and accordingly can serve as a basis for elemental
identification, similar to the applications
of XPS and AES. In theory, EELS should
be a good technique for the characterization of organic materials since inelastic
electron scattering is especially sensitive to
low mass elements such as carbon, nitrogen and oxygen, which prevail in polymeric materials (Williams, 1984). However,
because of the high energies of the input
electrons used and the beam currents necessary to obtain signals, serious damage
occurs and thus only limited applications
of EELS has been seen for the analysis of
polymeric materials. Further, the sample
preparation for EELS analysis in the
transmission mode is very tedious, since
extremely thin films have to be prepared;
the specimen thickness should be significantly less than the inelastic mean free
path length, which has been measured to
be about 160 nm for organic polymeric
materials at the energies used (Reimer,
1984). One should note the similarity of
the conditions for EELS and TEM, viz,
>10 2 KeV electron beams. On this basis
EELS offers the possibility of very good
lateral resolution, on the order of a few

21.2 Survey of Characterization Techniques

nanometers, for materials analysis (Briber


and Khoury, 1988; Cazaux and Colliex,
1990).
The EELS spectra of several polymers
have been published in the literature
(Ritsko, 1979; Ritsko and Bigelow, 1978;
Ritsko etal., 1978): polyethylene, polystyrene, polyvinylpyridine and polyoxymethylene. However, no quantitative
analysis was conducted to determine the
relative compositions. Briber and Khoury
(1988) investigated the possibility of using
EELS to measure quantitatively the atomic compositions of polymeric materials on
a local scale using Nylon-66 and poly
(chlorotrifluoro ethylene) as model systems. Reasonable agreement between the
measured and calculated compositions of
the two polymers was obtained. Special
precaution was taken such that no significant radiation of the sample occurred during the measurements. Such measurements
have shown that EELS is a tool for examining the composition of polymers on a
scale of microns or less if proper precautions are taken to evaluate the radiation
damage effects on the spectra prior to the
final interpretation of the data. It has also
been shown that this technique has the potential for the analysis of the composition
of different phases in polymer blend systems (Briber and Khoury, 1988).
21.2.4.3 High Resolution Electron Energy
Loss Spectroscopy (HREELS)
The HREELS experiment is essentially
the same as an EELS experiment except
that much lower energy electrons are used
to scatter off the sample and much smaller
energy losses are measured with very high
precision. The measurements are directed
at determining vibrational modes of the
surface of a material and thus should be
ideal for organic materials, both molecules

713

and polymers, which exhibit rich vibrational structure associated with the variety
of chemical bonds and symmetries in the
molecules. For this reason, HREELS is
now being developed rapidly into one of
the more useful techniques for probing
physical and chemical phenomena at surfaces (Erskine, 1986). For optimal excitation of the surface vibrations, low energy
electrons, typically 1-10 eV, are used. In
this energy range the mean free path of the
electrons is quite short, 2 nm, and therefore transmission experiments are nearly
impossible since free standing films this
thin are rarely available. Consequently only reflection experiments are done. The energy losses of interest range between zero
and a few hundred meV and correspond to
excitations of both localized vibrations of
molecular groups and of intrinsic collective vibrations (phonons) of any regular
lattice structures present (Pireaux et al.,
1990 a, b). For a detailed discussion of the
interaction mechanisms and analysis aspects, the reader is referred to recent reviews and a book on HREELS (Pireaux
etal., 1990a, 1991; Dubois, 1993; Ibach
and Mills, 1982).
There are several distinct advantages of
using HREELS to characterize the surface
of a material. One of the most important
features of HREELS technique is the high
surface selectivity which arises because of
the short mean free path of the low energy
electrons; typically the uppermost 2 nm of
an organic material (Gardella and Pireaux,
1990). Like other vibrational spectroscopic
techniques, HREELS is sensitive to the
molecular nature of the analyzed material.
In particular, it is possible in principle to
obtain chemical structure, tacticity and
orientational information about groups in
the surface region (Pireaux et al., 1991). Of
particular advantage in surface analysis, is
the fact that species including hydrogen

714

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

can be probed in HREELS, a very difficult


task for most other surface analysis methods which are sensitive to virtually every
element in the periodic table but hydrogen,
e.g., XPS and AES. In HREELS, H is indirectly detected through its influence on
molecular vibrations, e.g., of CH^, O - H
and N - H x functionalities in organic materials (Pireaux etal., 1990b). Under ideal
conditions, HREELS spectrometers are
capable of detecting surface layers of
molecules at concentrations below 10" 3
monolayers (Erskine, 1986).
There are also disadvantages associated
with HREELS analysis. Of particular concern for organic materials is the possibility
of electron induced damage. However,
since the energy of a typical electron beam
is of the order of 10 eV, no sample damage
is expected unless, as will be discussed below, an electron flood gun is used to neutralize the charged surfaces in nonconducting materials such as polymers. HREELS
can be viewed as the surface counterpart of
the classical photon in-photon out infrared and Raman spectroscopies (Pireaux
et al., 1990 b), since they essentially deliver
the same types of physical information
(Ibach, 1977). However, a distinct difference is that traditionally, HREELS has
suffered from much worse instrumental
resolution in comparison with these photon spectroscopies. For example, with infrared spectroscopy resolutions of fractions of an meV are easy to achieve whereas the resolution of current HREELS instruments are in the range of 3-10 meV, as
measured by the FWHM (full width at half
maximum) of the elastic (zero energy loss)
peak. Recent development has shown that
spectral resolution (FWHM) better than
1 meV can be achieved in HREELS
(Willis, 1990), which is approaching that in
infrared reflection absorption spectroscopy (see Sec. 21.2.2.8). However, such reso-

lution is difficult to come by and is not


present in standard instruments. As would
be expected for an electron in-electron
out technique, another problem for many
organic materials is the surface charging
associated with the analysis of nonconducting samples. In fact, the charging
problem has been the main road block for
the application of HREELS to nonconducting materials. It has been found that
the charging problem can be resolved by a
simultaneous flooding of the whole sample
by a low current (< 1 mA) of high energy
electrons (l-2keV) (Liehr etal., 1986).
The use of an electron flood gun at these
energies, however, introduces additional
problems of electron-induced degradation
for organic materials (Pireaux et al.,
1990 a). Some finesse can be applied here
by using very thin (^15 nm) polymer films
deposited on a conducting substrate, a
strategy which can prevent surface charging and the associated degradation from
flood guns by quickly leaking off the accumulated surface charge to the substrate
(Vilar et al., 1987). A final complication is
that several types of scattering mechanisms involving interactions of the electron
energy and momentum with vibrational
states are known but it is often unclear
which one or combinations of them are
operative in a particular experiment. Since
the interpretation of the spectra in terms of
both structure and quantification require
the specification of an exact scattering
mechanism, until now it has not been possible to use spectral intensities from organic materials in a quantitative manner
(Pireaux et al., 1990a, 1991). The presence
of surface roughness of the scale of the
scattering depth, also provides a problem
with repect to entangling mechanisms
since one way of distinguishing mechanisms is to do careful measurements of the
angular dependence of the spectra on the

21.2 Survey of Characterization Techniques

angle of the scattered (output) beam; however, rough surfaces defeat these measurements. Groundwork still needs to be done
in order to understand the basic mechanisms involved in the scattering process in
HREELS.
The first reported application of HREELS to polymeric materials occurred in 1986
when the high resolution electron energy
loss spectrum of polyethylene was observed (Pireaux et al., 1986). Since then,
the applications have been extended to
other polymer systems, such as polystyrene (Vilar et al., 1987; Pireaux etal.,
1987; Vilar et al., 1989), polyimide
(Pireaux etal., 1990b, 1989), poly(methyl
methacrylate) (Pireaux et al., 1990a,
1991), poly(vinyl alcohol), poly(acrylic
acid) (Pireaux etal., 1989) and polycarbonate (Apai and McKenna, 1991). Several publications concerning the applications
of HREELS to the study of LangmuirBlodgett molecular films have appeared in
the literature (Wandass and Gardella,
1986, 1987). Recently, Purtell and Pomerantz (1991) have used HREELS to study
surfaces relevant to lubrication. Their particular example consisted of an adsorbed
layer of stearic acid on sputtered carbon.
Since HREELS is very sensitive to the
molecular nature of the species, the technique also has been used to investigate the
chemical interaction between a polymer
and a metal by vacuum depositing a thin
layer of metal onto a polymer surface. The
systems studied include Al-polyimide
(Pireaux etal., 1988, 1989, 1990b), A l poly(vinyl alcohol), Al-poly (acrylic acid)
and Al-polystyrene (Pireaux et al., 1989).
The reverse process, i.e., the deposition of
polyimide precursor onto a metal substrate, has also been employed to form
metal-polymer interface and the interface
then studied by HREELS (Jones et al.,
1989 a). The HREELS spectra of pure

715

polyimide and aluminum-coated polyimide are shown in Fig. 21-15 (Pireaux


etal., 1989). As shown in this figure, the
partial disappearance of the C = O stretching mode feature and the parallel intensity
reduction of the symmetric vO = C)
band indicate that the interaction of the
evaporated Al atoms involve the - C = O
site, and the simultaneous intensity decrease of the (C-N) band is consistent
with an influence of the deposited Al
atoms on the neighboring C - N bond.
Such results are quite complementary to
those obtained from XPS analysis on these
systems (see the previous section on XPS).
Recent applications to organic surfaces
have been reviewed (Dubois, 1993).
Loss energy (MeV)
200
400

1000
2000
3000
Wave number (cm"1)

Figure 21-15. High resolution electron energy loss


spectroscopic (HREELS) vibrational spectrum of the
interface formation between a polyimide film and
evaporated aluminum, (a) Clean polyimide film surface, (b) With 1/10 layer of aluminum, (c) With half a
layer of aluminum. Reprinted with permission from
Pireaux etal. (1989). Copyright Springer-Verlag
GmbH & Co.

716

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

The application of HREELS to polymers is relatively new in comparison with


the application of other surface analysis
techniques. The potential of this technique
has emerged in providing information
about the surface functionality and the orientation of such functionality with respect
to the surface plane of the material
(Pireaux etal., 1991). However, the realization of the true potential of the technique will rely on an improved fundamental understanding of the mechanism involved in the interaction of low energy
electrons with the surface of solid materials.
21.2.4.4 Inelastic Electron Tunneling
Spectroscopy (IETS)

The IETS technique is a very unique surface analysis experiment, unlike any other
type, in that it involves the fabrication of a
planar, solid state metal-insulator (includes sample)-metal (MIM) tunneling
device for each measurement and after the
experiment the device is discarded. In view
of this the only types of samples which can
be analyzed are adsorbed monolayers and
nanometer-scale thin films which can be
sandwiched between a conducting substrate with a nanometer-scale, pinhole free,
dielectric overlayer, usually Al with a native oxide film, and a vacuum-deposited
metal, top electrode capable of undergoing
a transition to a superconducting state at
experimentally reasonable temperatures;
usually Pb is chosen to serve this function.
This alone probably would have discouraged completely the use of this method but
for the fact that, very early in its development, IETS was shown to be an incredibly
sensitive technique for measuring the vibrational spectra of molecules adsorbed
on planar metal oxide surfaces, particularly A12O3, an oxide of practical interest

(Hansma, 1987). Further work has shown


the technique to be quite useful in some
specific instances for the analysis of the
bonding and chemical reactions between
organic molecules, including polymers,
and a few metal oxides, again especially
A12O3. Some practical applications have
been found such as the adhesion of paints
on a metal (oxide) surface (Henriksen
etal., 1988; Colletti etal., 1987).
The principle of IETS has been well documented in the literature (Hansma, 1987;
Henriksen etal., 1988; Lambe and Jacklevic, 1968) and will only be discussed
briefly here. When a voltage bias is applied
across the MIM device, electrons begin
tunneling across the insulating barrier and
create a current. When the energy of the
tunneling electrons is equal to (resonant
with) the vibrational energy of the adsorbed molecules on the substrate (metal
oxide) surface the tunneling current
changes and signals the presence of the vibrational state at an energy of the electron
determined by the voltage setting; this
spectra are obtained. An important point
is that both Raman and infrared active
transitions are allowed and so additional
spectral features are available for interpretation compared to either of the latter
spectroscopies. It has been shown that the
sensitivity of IETS is sufficient to see just a
small fraction of a monolayer of an organic molecule over a junction area of less
than 1 mm 2 (Hansma, 1987; Cederberg,
1981). However, the exact thickness and
character of the molecular layer is critical
since it has been shown that in some cases
even densely packed monolayers can give
rise to problems (Sondag et al., 1991) and
ordinarily one monolayer is best for a tunneling spectrum (Hansma, 1987). In order
to see typical vibrational features of interest the energy resolution must be on the
order of 1 meV or better (refer to the dis-

21.2 Survey of Characterization Techniques

cussions above for HREELS). This can be


achieved by the use of liquid He cooling
which serves both to reduce thermal
broadening effects in the electron energies
and to allow the top electrode to go superconducting which improves the device operation. The instrumental resolution of
typical tunneling spectra ranges from 1 to
4meV and the best resolution is about
2 cm" 1 . The devices can obtain useful
spectra from energies above the phonon
regions of the electrode metals ( ^ 1 0 0 200 cm" 1 ) into the energy regions of electronic transitions.
In the last twenty years IETS has been
used extensively in the analysis of the adsorption of small molecules on metal oxide
surfaces. For a detailed account of such
applications, the reader can refer to review
papers by Hansma (1987, 1977). Several
groups also have used IETS to study the
adsorption of polymers on aluminum and
magnesium oxides. The polymers include
poly(vinyl alcohol), poly(acrylamide),
poly(methyl methacrylate), poly(vinyl acetate), poly(acrylic acid) and thermosetting
polymers (Mallik etal., 1985; Colletti
etal., 1987; Comyn etal., 1981, 1983).
However, as can be concluded easily
from the above discussion, IETS suffers
from several serious drawbacks. The sample preparation is complicated and involves very specific steps with severe restrictions on materials. A schematic representation of an IETS sample is shown in
Fig. 21-16. The preparation sequence is:
the substrate metal, usually Al, is vacuum
deposited on a glass slide, the metal is oxidized to form the insulating barrier, the
material of interest is adsorbed or coated
onto the barrier oxide, the lead top electrode is vacuum deposited onto the adsorbed layer and finally, the two electrodes
are connected to the measurig circuit. For
measurement the device is immersed into

717

Polymer

Inelastic Scattering
Elastic Scattering
El

Metal 1

Metal 2

Figure 21-16. Schematic representation of typical


sample geometry used in inelastic electron tunneling
spectroscopic (IETS) measurements.

liquid He. Extreme care has to be exercised


in sample preparation in order to achieve
data reproducibility. Besides, IETS is restricted to the analysis of molecules adsorbed on a few metal oxide surfaces such
as oxidized aluminum and magnesium, although it is expected that IETS can be
extended to other systems in the future
(Hansma, 1987). All of these drawbacks
have significantly limited the application
of IETS in the study of the adsorption of
organic molecules on metal oxide surfaces
and have favored the use of other techniques for obtaining vibrational spectra
such as infrared reflection-absorption
spectroscopy which works quite well on Al
surfaces with oxide overlayers, for example (see Sec. 21.2.2.8).

718

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

21.2.4.5 Scanning Tunneling Microscopy


(STM) and Atomic Force Microscopy
(AFM)

The invention of STM and AFM


(Binnig et al., 1982, 1986) has opened new
frontiers in surface science and technology
(Magonov, 1993) because the two techniques allow one to see, for the first time,
combinations of topographic and electronic features, resolved at the atomic scale, in
the surfaces of bulk materials. Each of
these methods depends upon the movement and placement, controlled at subangstrom precision by piezoelectric transducers, of an atomically sharp tip adjacent
to a flat surface. The presence of features
on the surface can be sensed as a function
of vertical distance by either variation of
electron tunneling current for a combination of a voltage biased metallic tip and a
conducting substrate, the STM method, or
by the variation of tip-sample forces, the
AFM method. Transmission electron microscopy and diffraction, of course, have
provided atomic resolution information
for a number of years but only ultrathin
films supported on conducting grids are
amenable to analysis and sample preparations often are arduous. One major difference then between the TEM and the scanning tip probes is that a vast range of sample types are amenable to the latter as compared to the former. Primarily because of
the wide sample range possible, the scanning tip techniques, although quite new,
have seen an explosion of applications to
analysis of surfaces of many materials.
Perhaps of all the materials which can be
studied, organic materials may be the most
difficult for these techniques because of the
typical insulating properties, a problem for
STM and because of the high propensity
for mechanical deformation ("soft" materials), a problem for both AFM and STM.

However, applications to organic materials are ever increasing and, in view of the
fact that many groups are presently actively pursuing refinements in the detailed interpretations of the images as well as further improvements in the instrumentation,
it can be expected that STM and AFM will
soon become well established tools in surface analysis of organic materials. STM
and AFM are treated in extenso in Chap.
11 of this Volume.
As mentioned in the introduction to this
section, the STM technique is based on the
quantum-mechanical phenomenon of electron tunneling between a sharp metallic
tip, often tungsten, and contiguous surface
atoms. With respect to the tunneling phenomenon, it is identical to IETS but differs
in detail in that the source of the tunneling
in the latter is a macroscopically large surface plane rather than a localized, atomicdimension location. Further, in IETS the
top electrode touches the tunneling film
surface whereas in STM (or AFM) contact
is not necessary and usually not desirable
for reasons of mechanical damage and/or
elastic deformation. The level of the tunneling current measured at these selected
locations on the surface yields maps of the
electronic structure convoluted with the
physical topography of the surface. In
many cases a simple interpretation is possible and leads to a map of the physical topography. The AFM technique depends
upon the force, repulsive or attractive, between a tip, generally a very hard material
such as silicon nitride, and contiguous surface atoms. With AFM a potential energy
type of map of the surface is thus created
which in its simplest interpretation provides information about the physical topography of the surface. In order to fully
understand the sample characteristics
from STM and AFM images, the exact
details of the interaction mechanisms obvi-

21.2 Survey of Characterization Techniques

ously must be known but often they are


not without extensive study and supporting theory, which unfortunately is still in
its infancy in this area. At present, it seems
fair to say that STM is more advanced in
both experiment and theory than AFM. A
detailed discussion of the principles involved in these techniques is beyond the
scope of this review and the interested
reader is referred to many reviews and
books that specifically deal with STM and
AFM (Magonov, 1993; Frommer 1992;
Sarid and Elings, 1991; Sarid, 1992; Guntherodt and Wiesendanger, 1992). Simple
schematic diagrams of the probe-sample
configurations of STM and AFM are given in Fig. 21-17.
In selecting an scanning tip imaging approach to analyzing the surfaces of organic
structures, one must be very mindful of the
particular problems associated with these
materials. The STM technique has remained quite limited in its applications because of the requirement of the material to
be electrically conducting to at least the
degree required to pass the very small tunneling currents. One straightforward application has been in the analysis of conducting polymers such as polypyrrole and
polyaniline (Madsen etal., 1991; Magonov etal., 1993 a). One type of organic
sample amenable to STM analysis is a thin
film supported on a smooth, ordered and

sample
surface

STM

AFM

719

conducting substrate, e.g., highly oriented


pyrolytic graphite (HOPG) or oriented,
crystalline gold films deposited on cleaved
mica, in which the organic film is sufficiently thin to allow a measureable tunneling current. An example of the atomic resolution, topographical detail which can be
provided for such samples is given in
Fig. 21-18, which shows an image for an
ordered film of a liquid crystalline molecule self-assembled on a graphite surface.
However, in spite of the conducting substrate and near-molecular film thicknesses,
these types of samples must be imaged
carefully since it is known that the use of
excessive bias voltage can result in improper imaging and damage to the surface
(Kim and Bard, 1992). In order to really
access surface characterization of a broad
spectrum of organic materials, the necessity of a conducting substrate and the restriction of extremely thin films must be
overcome. One possibility for this is
through a recent advance using microwave
frequency alternating current (AC) conditions between the tip and the sample which
allows insulators to be analyzed since only
transient polarization of the substrate is
required to obtain a tunneling current at
these frequencies (Stranick and Weiss,
1993). This AC technique also has the advantage that the tunneling current can be
measured as a function of the AC frequen-

Figure 21-17. Schematic representation of STM and AFM measurements.

720

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

Figure 21-18. (a) STM image of w-alkylcyanobiphenyl


(nCB) (ft = 10) molecules absorbed on graphite, (b)
Model showing the packing of the 10CB molecules.
The shaded and unshaded segments represent the alkyl tails and the cyanobiphenyl head groups, respectively. Reprinted with permission from Smith et al.
(1989). Copyright The American Association for the
Advancement of Science.

cy when tunable microwave circuits are


used and very molecule-specific dielectric
spectra can be generated (Stranick et al.,
1993). This technique has the promise of
being the first chemically sensitive atomic
resolution imaging probe if unique fingerprint type spectra can be generated for
specific functional groups.
With the exception of the newly developing AC-STM technique, only AFM can
be used for imaging nonconducting materials as it is not sensitive to the electrical
properties of the material. For this reason,

applications to the analysis of molecular


and polymeric organic materials have been
expanding dramatically in the last few
years and recent reviews are available to
find many examples (Magonov, 1993;
Snyder and White, 1992; Frommer, 1992;
Reneker, 1990). Applications have included the imaging of crystalline ordering at
polymer surfaces (Stocker etal., 1992;
Lotz etal., 1991) and the analysis of the
orientation of polymer chains under different processing conditions (Magonov et al.,
1993 b). Recent experience has shown that
due caution must be exercized in interpreting both STM and AFM images since improper imaging techniques may result in
the introduction of artifacts into the images (Nawaz et al., 1992; Leung and Goh,
1992). Leung and Goh (1992), for example, have demonstrated that by applying
excessive force (in the vicinity of 10 ~7 N)
in AFM measurements, the interaction between the metallic tip and polystyrene film
can result in the formation of a tip-induced, ordered structure on the surface of
polystyrene. It has been suggested that the
operating force for studying biological
samples should not exceed 1 0 - 1 1 N and
that forces of the order of 10 ~8 N can result in large deformations of the surface
(Abraham and Batra, 1989).
21.2.4.6 Low Energy Electron Diffraction
(LEED) and Transmission Electron
Diffraction (TED)

Electrons can be accelerated to energies


which impart corresponding deBroglie
wavelengths to them of the order of atomic
spacings. Under these conditions diffraction can occur and the electron thus can be
used as a structural probe of ordering. At
relatively low energies the electrons also
exhibit the added advantage that the diffractive scattering from a surface effective-

21.2 Survey of Characterization Techniques

ly probes only the outermost atomic layers


of the target material. This has given rise
to the LEED technique in which electrons
with low energies ranging from 20 to
300 eV are reflected off a surface. The
mean free path for inelastic scattering of
electrons in this energy range is typically
about 1.0 nm, a minimum compared to the
values for all other electron energies (Seah
and Dench, 1979). Thus, LEED is extremely surface selective and this attribute
has helped to make the technique a very
popular method for characterization of ordering in single crystal inorganic surfaces
of metals and semiconductors (Woodruff
and Delchar, 1986). However, application
to organic films has been relatively rare
until recently because of problems with
sample charging and damage caused by
the high electron flux necessary for the
generation of good quality diffraction
spots. In order to counter the charging
problems expected with electron bombardment LEED studies are restricted to
monolayer films of organics on metal substrates. A related technique is TED which
relies on the transmission of an electron
beam through a thin (tens of nanometers
or less) conductive film, rather than reflection as with LEED. In order to permit
transmission of the electron beam through
a material, much higher energies are necessary than for LEED. Such high energies
are conveniently available in an electron
microscope (~ 102 keV) and so an advantage of the TED experiment is that the
sample also can be imaged at high resolution. Again, because of sample damage
and charging effects this experiment is difficult for organic materials but several excellent examples of applications have been
reported.
The early attempts to use LEED for organics were unsuccessful because of electron beam damage but studies of electron-

721

damage stable adsorbates showed the possibilities (Firment and Somorjai, 1979) and
improvements in analysis methods and instrumentation prompted other studies
(Ohtani etal., 1986; Bent and Somorjai,
1989). Of particular interest are polyatomic adsorbates and examples now exist
of a variety of molecular films. Lipid-like
films have been analyzed (Vogel and Woll,
1988). Gui et al. (1991) have examined adsorbates such as thiophenol, benzyl mercaptan and alkyl mercaptan on Ag(lll)
and Pt(lll) electrodes and Dubois etal.
(1993) have carried out analysis of self-assembled monolayers formed from the
chemisorption of /7-alkane thiols on gold.
The latter studies show that the "probing
depth" of LEED in organic materials can
be less than 2.5 nm, a conclusion based on
the fact that the substrate (gold) electron
diffraction spots were not observed even
though the thickness of the self-assembled
monolayers analyzed is only about 2.5 nm
(Laibinis et al., 1991). While very valuable
information has been obtained in these
studies about the ordering and structure of
the organic films on metal substrates, one
however, should use caution in applying
such analyses to organic materials because
of potential damage upon exposure to electron beams. Dubois et al. (1993) have observed that even when very low electron
fluxes are used (below 100 pA/mm 2 , approximately four orders of magnitude less
than in a conventional LEED experiment)
significant damage can occur to the organic overlayers after only a few minutes.
The TED technique is an alternative to
LEED for cases in which the organic film
of interest can be supported on the surface
of a conducting film which is sufficiently
thin to allow enough transmission of an
electron beam to provide the analysis.
Garoff and Deckman (1986) have characterized the structure of Langmuir-Blod-

722

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

gett monolayer films on conductive amorphous carbon films. Dorset (1990) has
characterized ultrathin films of crystalline
alkanes on benzoic acid substrates. In
these cases the ordering is not dependent
on the substrate so the use of vacuumdeposited carbon provides a convenient,
transmissive substrate. Ordered films
which require ordered substrates, such as
metal single crystals are a problem since
metals are much less transmissive than
amorphous carbon and few examples of
nanometer scale thick single crystals exist.
A good example of a tractable system is
that of ordered self-assembled monolayers
of alkanethiolates on thin Au(lll) films
produced by vacuum deposition methods
(Strong and Whitesides, 1988). This study
was the first to show that these monolayers
are hexagonally arranged on the Au(lll)
surface.
Both LEED and TED offer excellent
ways to characterize the ordering in thin
organic films when suitably conducting,
non-interfering substrates are used and
when the experiments can be done extremely carefully at low beam currents to
avoid sample damage. More examples are
expected to be seen in the future.
21.2.5 Ions or Neutrals In, Ions
or Neutrals Out
Ions have been one of the traditional
surface probes over the years. Because
ions, like electrons, are relatively easy to
manipulate by well-understood principles
of electrostatic (and magnetic) field configurations, ion spectroscopies have evolved
synchronously with incremental increases
in the ability to engineer sophisticated generation and analysis systems of charged
particles. Ion beam generation can be as
simple as an inexpensive add-on ion gun
for an existing vacuum chamber to give

keV energy beams or as complicated as the


installation of a high energy ion accelerator system to give MeV beams of heavy
ions. Because of the dominance of Coulombic over other types of forces in particle interactions, at first glance it may appear that all charged particles might behave similarly upon impacting solids and
thus that both electrons and ions should
have similar surface spectroscopies. However, they are quite different. The major
difference between electrons and ions as
input probes is that ions can be generated
with much higher momenta than electrons
for the reasons that ions have orders of
magnitude heavier masses than electrons
and can be prepared with multiple charges.
The first reason is important because, for
constant kinetic energy, momentum increases with (mass)1/2 and thus higher
mass particles possess higher momenta.
The second reason is important because,
for a constant electric field gradient, the
acceleration of a charged particle increases
in proportion to the charge and thus
higher ionic charge leads to higher kinetic
energies. Relevant to the latter point is the
fact that after the creation of a beam of
free ions, the sign and number of the
charge on the ions often can be varied
rather easily by charge exchange processes
with the exact charge states possible dictated by the particular atoms involved. Further, one of the possible charge states is
zero and accordingly neutral particles with
controlled energy and momentum also can
be produced from an ion beam. The major
effect of increased initial momentum of a
probe particle, neutral or charged, is the
corresponding increase in momentum
transfer to the target solid during collision.
Thus it is understandable that almost all
the ion techniques depend upon momentum transfer effects as the critical aspect of
the analysis mechanism. The magnitudes

21.2 Survey of Characterization Techniques

of the velocities are important because


these regulate the residence time of the ions
in a given vicinity of the solid and thus
cause different effects depending upon the
time scales of different response mechanisms in the solid, e.g., polarization and
excitation of bound electrons. The magnitudes of kinetic energies are important
since they will govern the ability of local
attractive (or repulsive) potentials in the
solid to affect the particle trajectory in the
vicinity of the potential. Thus by tuning
energy and mass, the exact mechanisms of
interaction can be altered in a useful way
to create analysis spectroscopies. These
mechanisms include: (1) completely elastic,
billard-ball like, collisions with target nuclei-these can include ejection of a target
matrix particle by an elastic recoil with the
input particle left behind in the solid;
(2) highly explosive collisions which irreversibly damage the solid in the vicinity
along the impact trajectory and eject particles from the surface and (3) resonant collisions which cause excitation of discrete
quantum processes. The most useful of the
latter type of collisions are those which
occur with specific nuclei at very high and
specific kinetic energies (MeV) to induce
nuclear reactions leading to ejection of a
new particle(s), either matter or photons.
Surprisingly, the elastic character of a collision does not show a monotonic dependence on the kinetic energy of the input
ion; very low (< several keV) and very high
(several MeV) energies can lead to elastic
collisions whereas the intermediate energies ( ^ 102 keV) can lead to inelastic collisions. The reasons for this are, however,
well known and depend upon the interplay
of the factors given above. From these
mechanisms three types of spectroscopies
arise. First, for ions (or neutrals) with intermediate energies and momenta which
give rise to inelastic collisions followed by

723

ejection of different fragments from the


target matrix, output particle energies are
meaningless in terms of analysis and only
the identity and number of the fragments
are of value (SIMS techniques). Second,
for ions which give rise to elastic collisions,
the energies, momenta and number of the
output particles is measured and used to
identify the number and types of atoms
present in the collision region of the target
solid (ISS, RBS, FRES). Third, for ions
with the correct energies to induce resonant quantum processes in a specific type
of nucleus, the output particle, which for
the nuclear reaction processes is different
than the input particle, is counted and used
to determine the number of the specific
resonant nuclei present in the collision region (NRA). For the high energy spectroscopies, the probe particle penetrates
deeply into the target solid with a gradual
loss of kinetic energy. This energy-depth
dependence can be used to provide analyses of the constituent atoms of a solid as a
function of depth. Finally, the destructive,
inelastic types of collision processes can be
used to advantage as a means of removing
or sputtering away sequential layers of solid to continually expose new regions of the
material. When this sputtering is combined
with a highly surface selective probe, a
depth profile analysis of the material can
be made. This aspect will be discussed in
another section.
21.2.5.1 Rutherford Backscattering (RBS)

There are a group of surface analysis


techniques that depend upon the use of
very energetic ion beams with energies in
the range of 1 - 3 MeV. Of these, which include RBS, FRES and NRA, RBS is the
most popular and the most generally
useful one. It can be used for elemental
identification and nondestructive (usually)

724

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

depth profile analysis of almost any solid


including many organic materials. The
principle of RBS is surprisingly simple due
to a fortuitous combination of factors involving the response of the solid matrix to
the passing of an extremely energetic ion,
usually 4 He 2 + . Essentially, the ion can be
considered as a featureless projectile which
is moving rapidly through a slightly viscous medium (electron clouds) with widely
separated, small scattering centers (nuclei).
As the ion moves through the medium it
slowly decelerates due to the viscous drag
(stopping power) of the electrons. This
drag amounts to a myriad of inelastic collisions with the electrons in the solid with
dissipation of kinetic energy into electronic
excitations (Feldman and Mayer, 1986).
The energy lost in penetration through the
solid is directly proportional to the path
length. Eventually the ion may strike a

(a)

Nuclear Particle Detector

nucleus and recoil via a momentum exchange, decelerating suddenly. In a fraction of these events the ions will reverse
direction, decelerate slowly and then exit
from the surface. The energies of these
backscattered ions can be measured and
compared to the initial energy of the incoming ions thus generating energy loss
spectra. By applying the laws of momentum and energy conservation to account
for the large portion of the energy loss,
together with known values of the stopping power of the solid matrix to account
for the small part of the energy loss, the
atomic weight of the scattering nucleus can
be identified and the concentrations determined as a function of depth from the surface where the input ion entered. The RBS
experiment and related mechanisms are
schematically illustrated in Fig. 21-19. The
collision part of the analysis model given
above can be described by the equation
E

Scattered
Beam

(21-15)

Sample
Collimators

(b)

Ml

Figure 21-19. Schematic representations of (b) the


mechanism involved in the backscattering process
and (a) the typical experimental set-up in Rutherford
backscattering (RBS) spectrometry.

where Mx and Eo are the mass and kinetic


energy of the incoming ion, respectively,
M2 is the mass of the target nucleus initially at rest and 6 is the angle of backscattering, usually close to 180. All samples do
not lead to ideal analyses because of mitigating factors, e.g., the presence of one element may interfere with identification of
another. After penetrations of the order of
micrometers, the energy loss becomes sufficiently great that the ability to analyze
the data quantitatively becomes unreliable. For a detailed description of the RBS
technique, the reader is referred to recent
reviews and books on RBS (Feldman and
Mayer, 1986; Green and Doyle, 1990; Chu
et al., 1978; Bird and Willams, 1989; Boerma, 1990; Gossmann and Feldman, 1987).

21.2 Survey of Characterization Techniques

The techniques treated here and in the following two sections are dealt with in more
general terms in Chap. 17 of this Volume.
It is seen from the above discussion that
one advantage of RBS is the simplicity of
interpretation and the absolute, quantitative nature of the analysis, involving minimum calibrations. The depth profiling resolution is governed in ideal cases mostly by
the energy resolution of the experiment
and leads to typical depth resolutions of
^ 3 0 n m in polymers. This number effectively defines the surface selectivity. The
sensitivity of the method, in ideal cases of
heavy elements in a light element matrix is
fractions of a monolayer. Several elements
can be quantified simultaneously although
the presence of heavy elements often renders quantification of lighter elements
more difficult because of spectral interference. Initially, it was not imagined that
polymers could be analyzed by such high
energy particles without enormous damage and poor results. However, in an early
study of the migration of copper ions in
polyethylene undergoing thermal oxidation in contact with copper surfaces, it was
shown that the advantages of RBS could
apply to polymers (Allara and White,
1978; Allara etal., 1976). Since that time
the number of useful studies has been
steadily increasing. RBS has been used by
many researchers to study a variety of
problems, including surface modification
of polymers (Davenas etal. 1988, 1989;
Matienzo etal., 1988; Venkatesan etal.,
1983), diffusion of small atoms or
molecules into polymers (Mills et al., 1986;
Lasky et al., 1988), interdiffusion at polymer-polymer interfaces
(Rafailovich
etal., 1988; Shull etal., 1988) and diffusion at metal-polymer interfaces (Ohuchi
and Freilich, 1988; Chauvin etal., 1987;
Koh etal., 1990; Shanker and MacDonald, 1987; Green and Berger, 1993). The

725

development of in-air or external Rutherford backscattering spectrometry (X-RBS)


should make this technique more useful to
analyze practical problems, since in-situ
measurements of the diffusion processes
should be possible with X-RBS (Doyle
et al., 1991). Finally, the lateral resolution
capabilites should be mentioned. Although the ion beam spot on the sample
can be made to be less than one jam, the
constraints of the RBS experiment limit
the useful spot size to about 10 |im, obtainable with commercial instrumentation.
Little work seems to have been done with
small spot RBS on polymers.
As has been alluded to briefly above,
there are some disadvantages associated
with the application of RBS in analysis of
organic materials. First of all, RBS is relatively inefficient for detecting low mass elements because of interference from high
mass spectral contributions and low cross
sections for collision, and this has significantly limited its applications in polymeric
material analysis since low mass elements
prevail in organic materials. In fact, it is
not possible to detect H by RBS and this is
the prevalent element in most polymers.
Another ever-present problem with ion
beam analysis of organic materials is degradation. It has always seemed puzzling
that polymers could be analyzed accurately the high energy ions used in RBS and, in
fact, sample damage can be seen visibly as
dark spots the size of the ion beam even
though correct analyses are obtained. Actually, most polymers are radiation sensitive and with some polymers RBS analyses
can be affected by radiation damage, particularly if high doses are used. Namavar
and Budnick (1987) have investigated the
feasibility of using RBS to study polymers
in terms of beam damage during RBS measurements. They have found that for
poly(ethylene terephthalate) there is no de-

726

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

tectable compositional change as long as


the dose applied is lower than 5 x 10 14
4
He + + /cm 2 at room temperature and up
to 2 x 1015 4 He + + /cm 2 when liquid nitrogen is used to cool the sample. Valenty
et al. (1984) have performed detailed studies of the craters created during RBS ion
bombardment of a polycarbonate surface.
Using a combination of techniques including TEM, SEM, XPS, IRS and profilometry, they concluded that the crater consists
of densified material which results from
a combination of ion beam-induced intramolecular rearrangement, decarbonylation and free radical crosslinking of the
polymer. It is therefore important that care
be taken during RBS measurements on
new or unknown materials to ensure that
ion beam induced changes do not change
the elemental composition of the sample
when RBS is applied.
21.2.5.2 Forward Recoil Elastic Scattering
(FRES)
One of the problems with RBS mentioned above is the increased difficulty of
analysis of light elements, a particular concern for organic materials which primarily
consist of C, H, and O. This capability can
be achieved by a simple modification of
the RBS type of experiment from a
backscattering to a forward recoil measurement. Upon collision of an incoming
energetic ion with the nucleus of a light
atom in the target matrix, under the right
conditions of approach angle and projectile momentum, the light atom can be
knocked out of its position in the solid and
ejected from the surface. At the right energies this recoil process is elastic and the
RBS analysis methods apply. Thus by
measuring the recoil energies, the identities
of the atoms and their depth dependent
concentrations can be determined. The

special niche for FRES is the determination of hydrogen. The basic principles of
the technique have been well documented
in the literature and only a brief description will be made here. For a detailed discussion, the reader can refer to a recent
review by Green and Doyle (1990) and a
book by Feldman and Mayer (1986).
The schematic of a FRES experiment
and a typical spectrum are shown in
Fig. 21-20. In the measurement, a high energy beam with energy Eo (~MeV) impinges on a target at an angle a with respect to the surface plane of the sample.
The incident ions undergo a numbers of
kinematic collisions with the nuclei in the
target which results in the recoil of these
target nuclei at an angle 9. Forward scattering results when 6 < 90. An energy sensitive detector is placed at a fixed angle to
detect those nuclei which have energies
characteristic of their masses and of the
depth from which they recoil. For analysis
of H and D, a thin mylar foil is placed
directly in front of the detector in order to
block the penetration of the abundantly
scattered helium ions while permitting the
passage of the H and D ions to reach the
detector (Feldman and Mayer, 1986).
There are several advantages of using
FRES to obtain depth profile information.
First, it is a nondestructive technique in
the sense that no ion sputtering and material removal is necessary. However, as with
RBS, care should be taken in FRES measurements of organic materials since radiation damage may occur with the high energy ion beams used. Second, FRES experiments are relatively easy to do and data
acquisition times is relatively short. Third,
the quantification of FRES data is relatively easy since the collision process involved in FRES obeys the same principles
as in Rutherford backscattering spectroscopy (RBS) which is well understood and

21.2 Survey of Characterization Techniques

727

10 (im thick
mylar stopper
foil
*He2+ ions
Energy
sensitive
(a)

detector
1.0

1.2

Energy (MeV)
U

1.6

Ad-PS(255K) on PS(20M) unannealed


d-PS(255K) on PS(20M) annealed 3600 s at 170 C

Figure 21-20. (a) Schematic representation of the geometry of forward recoil spectrometry (FRES).
(b) Two FRES spectra, one for a
thin layer (~40.0 nm) of deuterated polystyrene (d-PS) on top of
a thick layer of PS (triangles) before annealing and one for the
same bilayer after annealing at
170C showing the effects of diffusion of d-PS below the surface.
Reprinted with permission from
Mills et al. (1984). Copyright
American Institute of Physics.

4
A

T5

1u
"surface

200
(b)

250

300

350

Channel

well developed. However, in comparison


with other non-destructive depth profiling
techniques such as RBS and NRA, the
depth resolution of FRES is relatively
poor under standard analysis conditions,
i.e., with a 4 He + + ion beam, the one which
is the most often used for polymer analysis, and with the use of conventional silicon barrier detectors. The poor depth resolution in conventional FRES arises from
the energy straggling of the recoiled ions
when passing through the mylar foil in
front of the detector. In addition to degrading the depth resolution, the absorber
foil also limits the accessible depth range of
depth profiling information (Green and
Doyle, 1990). Recently, the development

of a time-of-flight spectrometer has successfully led to the replacement of the absorber foil. With a time-of-flight tube, a
depth resolution of about 30 nm has been
achieved (Sokolov et al., 1989; Rafailovich
et al., 1990), and even better depth resolution (about 20 nm) has been reported in
the literature (Stamm, 1992). Besides using
4
He + + as the probing beam in FRES, with
the limitation that only hydrogen and deuterium can be detected, beams of heavier
ions, such as Si, S, and Cl, can be used as
the incident probes. Their use enables the
detection of other light elements such as C,
N, and O, commonly found in polymers
(Green and Doyle, 1990; Baglin and
Willams, 1989; Gujrathi, 1990). Higher

728

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

depth resolution has been achieved by using such heavier ion beams. Groleau et al.
have estimated that the depth resolution
for light elements is about 8-9 nm at the
surface in their FRES work using 30 MeV
35 Cl ions as beam probe (Currie et al.,
1984; Groleau et al., 1983).
Although the application of FRES to
the study of polymers started less than 10
years ago (Groleau et al., 1983; Mills et al.,
1984), it has been used extensively to investigate a variety of problems such as the
diffusion of small molecules into polymers
(Mills et al., 1984), interdiffusion of identical or different polymers (Sokolov et al.,
1989; Gall and Kramer, 1991; Jones et al.,
1989; Composto etal., 1990; Shull et al.,
1991a), segregation of block copolymers
into the interface of two homopolymers
(Green and Russell, 1991; Shull etal.,
1991 b, c) and the interdifussion in metalpolymer interfaces (Currie etal., 1984;
Groleau et al., 1983; Chauvin et al., 1987).
A useful example of the application of
FRES to polymers can be given by the
study of the segregation of diblock copolymers into the interface of two homopolymers (Shull etal., 1991b, c; Green and
Russell, 1991). This phenomenon is important in practical applications of polymer
blends. It has been observed that adding
small quantities of an appropriate diblock
copolymer to a blend of immiscible homopolymers which have affinities for separate
blocks of the copolymer, can result in
high-performance plastics with appreciably improved mechanical properties. It
was proposed that the copolymer chains
may segregate to the interface between the
two homopolymers, thus reducing the interfacial tension and enhancing the mechanical strength of the interface (Shull
etal., 1990; Deline etal., 1991; Russell,
1991). However, there was no direct quantitative experimental evidence for such seg-

regation at the interface prior to the development of FRES, since few techniques
have the unique ability to selectively probe
depth regions of a polymer. Two block copolymer/homopolymer systems have been
studied so far. One is the diblock copolymer of polystyrene (PS) and poly(2-vinylpyridine) (PVP) in PS and PVP homopolymers and the other is the diblock copolymer of polystyrene and poly(methyl
methacrylate) (PMMA) in PS and PMMA
homopolymers. Shull et al. (1990) have investigated the segregation of unsymmetrical diblock copolymers of PS and PVP into
the interface between the two homopolymers, PS and PVP, while Green and Russell (1991) have studied the segregation of
the symmetric diblock copolymer of PS
and PMMA to the interface between the
two homopolymers, PS and PMMA. The
enrichment of the copolymer at the interface of the two homopolymers, PS and
PVP, is shown in Fig. 21-21 (after Shull
et al. 1990). The enrichment of the diblock
copolymer at the interface is compared

0.10

0
-2000

2000 4000
Depth (A)

6000

8000

Figure 21-21. Unsymmetrical diblock copolymer


(PS-PVP) distribution in the two homopolymers after annealing for 8 hours at the temperature of 178 C.
Reprinted with permission from Shull et al. (1990).
Copyright American Chemical Society.

21.2 Survey of Characterization Techniques

with the initial concentration profile without annealing as shown in Fig. 21-22 and
the initial sample geometry in Fig. 21-23.
At present the major limitation of FRES
appears to be the lack of depth resolution
necessary for many of the interesting problems in polymers, particularly interdiffusion between different phases. Since the
0.20

729

depth resolution is normally about 80 nm


when a 4 He 2 + ion beam is used, FRES has
been used in experiments involving long
diffusion times (Russell, 1991) which allow
the molecules to move many times their
contour lengths and thus give profiles
which fall within the resolution range of
the analysis. This approach unfortunately
has significantly limited the information
available about the initial stage of the interdiffusion process between polymers.
However, with the improved depth resolution from time-of-flight detectors and other improvements, it can be expected that
FRES will be used more extensively in the
future for the study of polymer materials,
especially for interfacial analyses.
21.2.5.3 Nuclear Reaction Analysis (NRA)

-2000

2000 4000
Depth (A)

6000

8000

Figure 21-22. Copolymer distribution in a typical


constant concentration sample (PS-PVP) prior to
the annealing step. Reprinted with permission from
Shull et al. (1990). Copyright American Chemical Society.

PS + d-PS-: n-PVP Copolymer


(-4000 A)

Figure 21-23. Initial geometry of the sample from


which data shown in Fig. 21-22 were obtained.

In addition to FRES for analysis of light


elements in surface regions of organic materials, another very similar method, using
highly specific nuclear reactions, offers an
alternative approach which yields similar
information but in some cases with improved precision. The basic method involves the onset of a nuclear reaction when
an incoming high energy ion (several MeV
range) of exactly the correct energy collides with a specific type of nucleus. The
nuclear reaction can produce a variety of
neutral or charged primary particles and/
or a photon (y-ray), dependig upon the
exact nature of the reaction. Although the
application of NRA to organic materials
is quite recent (Chaturvedi et aL, 1989;
Payne et al., 1989), it has been a traditional
technique for inorganic materials. For example, Lanford et al. have used NRA extensively in the study of hydrogen concentration depth profiles in a variety of inorganic materials since 1976 (Lanford, 1978;
Lanford etal., 1976; Pfeffer etal., 1982;
Schnatter etal., 1988; Krieger and Lan-

730

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

ford, 1988; Xie et al., 1988). There are a


number of different types of experiments
which can be done corresponding to a
number of different types of appropriate
nuclear reactions and to the different ways
of inducing and monitoring the reactions.
For organic materials the main interest has
been in analysis involving hydrogen and
the corresponding reactions have been carried out as either energy analysis or resonance experiments. The difference between
the latter two lies in whether the energy of
the incident ion beam is kept fixed or varied. In the energy analysis method, the energy of the ion beam is fixed and the energy
spectrum of the particles emitted by the
nuclear reaction recorded. In the resonance method, the energy of the incident
ion beam is varied and the nuclear reaction
yield is followed as a function of the energy
of the incident ion beam. These two methods are discussed in more detail below.
In the energy analysis method, the following nuclear reaction is the one most
used in polymer analysis:
3

He + 2 H -> 4 He +
Q = 18.352 MeV

(21-16)

where a 700 keV 3 He beam impinges on


the top of the polymer surface and either
the emitted energetic a particles or the proton is detected. In this approach, only the
concentration profiles of deuterium can be
measured since the reaction only occurs
for this isotope. Klein and co-workers
have used this reaction, by detecting the
emitted oc particle, to study the interdiffusion of polymers and have demonstrated
that a very high depth resolution can be
obtained with this technique (Klein, 1991;
Steiner et al., 1990; Chaturvedi et al.,
1990). It was shown that a depth resolution of 14 nm near the surface and 30 nm
at a depth of 100 nm can be achieved in

deuterated polystyrene (d-PS) (Chaturvedi


et al., 1990). The high depth resolution of
this method is demonstrated by the depth
profile of a 13.2 nm film of deuterated
polystyrene on a polished silicon substrate,
as shown in Fig. 21-23. Payne et al. (1989),
on the other hand, have detected the emitted protons and achieved a depth resolution of 30 nm at a depth up to 0.2 jim.
Such depth resolutions are higher than
those achievable from FRES (Jones et al.,
1989). Examples of applications to polymers using the energy analysis method include diffusion and mixing studies of different polymers: poly(vinyl chloride)polycaprolactone (Klein, 1991) and
deuterated polystyrene (d-PS) - hydrogenated polystyrene(h-PS) (Chaturvedi
et al., 1989; Payne et al., 1989; Klein, 1991,
Steiner etal., 1990; Chaturvedi etal.,
1990).
Even higher depth resolution in polymers has been demonstrated with the resonance NRA technique (Stamm, 1992;
Endisch et al., 1992). In this method, the
following nuclear reaction is used for polymers:
15

N + XH -*

12

C + 4He-h 4.43 MeV y-ray


(21-17)

In this reaction the sample is bombarded


with a 15 N beam and the yield of the characteristic 4.43 MeV y-ray is detected. If the
sample is bombarded with 15 N at the resonance energy ( rcs = 6.385 MeV), the yield
of the characteristic y-ray is proportional
to the numbers of hydrogen on the surface
of the target. However, when the beam energy is raised above the resonance energy,
the surface hydrogen can no longer be detected, but as the 15 N slows down passing
through the solid, it reaches the resonance
energy at a particular depth. Hence, by
measuring the y-ray yield as a function of
the energy of the 15 N, the concentration of

731

21.2 Survey of Characterization Techniques

hydrogen as a function of depth can be


determined (Lanford, 1978). The high
depth resolution is due to the fact that the
resonance energy width is very narrow and
that no energy loss occurs when the emitted y-rays travel through the solid, eliminating the struggling effect that greatly degrades the depth resolution in other high
energy ion beam analysis techniques such
as FRES and RBS (Chu et al., 1978; Feldman and Mayer, 1986). The use of this
technique is clearly demonstrated by the
recent work of Stamm in the study of
interdiffusion of polymers (Stamm, 1992;
Endisch et al., 1992). The depth resolution
was measured as a remarkable 3.5 nm at
the surface and approximately 16 nm at a
depth of 200 nm (Endisch et al., 1992). The
relatively poor depth resolution deep inside the solid is mostly limited by the straggling effect experienced by the incident
15
N beam. The high depth resolution of
the resonance method is demonstrated by
the result of the hydrogen density profile
analysis of a thin layer (93.3 nm) of hydrogenated polystyrene supported on a silicon
wafer substrate as shown in Fig. 21-25.
The extremely high depth resolution is
shown by the sharp slope of the plot at
zero depth of the polystyrene layer.
With the excellent depth resolution,
non-destructive profiling capability and
high detection sensitivity (10~ 2 -10~ 3
atomic fraction), the NRA technique can
be expected to be used more extensively in
the future in the study of the interdiffusion
and mixing of polymers, polymer blends
and the interface analysis. The recent development of in-air or external nuclear reaction analysis (X-NRA) makes this technique even more attractive for polymer
analysis (Doyle etal., 1991). However,
caution should be exercised when using
NRA for polymer analysis since radiation
damage of polymeric materials is generally

Energy (MeV)
5.85
5.80

5.90

5.75
100

IL

O0.8
c
O

60 c

.
1 \
If \

a
CD

0.2
o
>

80

^0.6

20

0
-20

0
20
40
Depth (nm)

60

Figure 21-24. Energy spectrum corresponding to


measured volume-fraction vs. depth profile of a thin
deuterated polystyrene (d-PS) film on a polished silicon substrate. The dashed line is a representation of
the ellipsometrically determined 13.2-nm-thick film.
Reprinted with permission from Chaturvedi et al.
(1990). Copyright American Institute of Physics.

40
60
Depth (nm)

Figure 21-25. Hydrogen depth profile of a sample


[93.3 nm hydrogenated polystyrene (h-PS) on a silicon wafer surface]. The depth resolution is approximately 3.5 nm near the surface of the film and about
16 nm at the depth of 200 nm inside the film.
Reprinted with permission from Endisch et al. (1992).
Copyright Elsevier Science Publishers.

not negligible. Short data acquisition time


or cooling of the sample by liquid nitrogen
should be used in order to ensure that radiation damage does not influence experimental results (Stamm, 1992). One also
should be aware that only limited numbers

732

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

of elements can be detected by current nuclear reaction methods even though the
knowledge of available reactions is still expanding for the analysis of different elements (Bird, 1989). It should be noted that
only hydrogen and deuteron are detected
in the two nuclear reaction analysis techniques discussed above. In terms of measurement convenience, it is important to
realize that data acquisition times of 10 to
100 h are necessary to achieve high sensitivity in nuclear reaction measurements
and accordingly this reduces the number of
samples that can be analyzed, a significant
limitation for a surface analysis method
(Bird, 1989).
21.2.5.4 Ion Scattering Spectroscopy (ISS)
When the energies of probe ion beams
are reduced from MeV to keV, dramatic
differences in the interactions with solids
occur. Unlike energetic beams which can
penetrate on the order of microns into
solids yielding information over a range of
depths, low energy ions scatter almost predominantly from the surface layer and are
of use only for near surface analyses. This
attribute can be optimized and used to advantage for "first monolayer" analysis
(Feldman and Mayer, 1986). At these low
energies an electrostatic analyzer can be
used for accurate resolution of the energies
of the scattered ions which then allows
very sensitive analysis of the surface composition. A typical ISS instrument will allow analysis of the top 0.3 to 0.5 nm of a
sample surface (Feldman and Mayer,
1986; Gardella and Pireaux, 1990).
In spite of the different surface sensitivities, the basic principle of ISS is similar to
that of RBS and has been well documented
by several authors (Feldman and Mayer,
1986; Fauster, 1988; Niehus and Spitzl,
1991; O'Connor and MacDonald, 1989).

Basically, ISS involves an eleastic twobody collision between a noble gas (unreactive) ion beam and the sample surface.
The energy of the incoming probe ions are
selected and the energy loss after scattering
at a chosen angle is measured. Since the
collisions are eleastic, the conservation of
momentum and energy can be applied and
the energy and numbers of the scattered
ions then directly yield information about
the masses and surface concentrations of
the atoms present in the target surface. In
addition, ISS has been used in the structure study of materials by making use of
the so-called shadow cone (Fauster, 1988,
Niehus and Spitzl, 1991; O'Connor and
MacDonald, 1989). The shadow cone is a
region of the surface which is not struck by
the probe ions because of the action of a
repulsive potential between the incident
ions and a particular target atom which
effectively prevents access of the probe
ions to a region behind the target atom.
This effect often depends upon the azimuthal direction of the probe ion beam
and thus when applied to single crystal surfaces it can be used to determine subtleties
of the relative positions of different atoms
in the surface lattice. ISS has been used
quite extensively in the structural analysis
of metal and semiconductor materials.
However, only limited applications have
been made in the area of organic materials
(Hook and Gardella, 1987; Hook etal.,
1987 a, b, 1986; Gardella and Hercules,
1981). Gardella and co-workers have investigated the possibility of using ISS
shadow cone effects to obtain surface
structure in polymers. In this study they
measured ion intensity ratio for scattering
from the surface C and O atoms in three
stereoisomers of poly(methyl methacrylate): isotactic, syndiotactic and atactic.
Based on space-filling molecular models,
the differences in the C/O scattered ion

733

21.2 Survey of Characterization Techniques

intensity ratios were attributed to differences in the surface structure of the three
stereoisomers (Hook et al., 1986). Since
these three polymers have identical compositions, the differences in scattering
must be due to subtle differences in which
atoms protrude furtherest in the outermost
surface. Further evidence for the sensitivity of ISS to surface structure of polymers
is provided by a study of two isomers of
poly (vinyl pyridine): one with the nitrogen
atom present in the ortho position in the
aromatic ring (P2VP) and the other in the
para position (P4VP) (Hook and Gardella,
1987; Hook et al., 1987a, b). Even though
same amount of nitrogen is present in the
two polymers, no nitrogen was detected in
P2VP whereas a nitrogen signal was observed for P4VP as shown in Fig. 21-26.
This difference has been explained in terms
of the shielding of nitrogen from the incoming ions by the main backbone of the
P2VP polymer.
Several other studies have been conducted using ISS to address different kinds of
problems associated with organic materials including both polymers and LB films.
For publications on the applications of ISS
to the studies of polymer surfaces prior to
1982, the reader can be referred to a review
paper by Baun (1982). ISS has been used in
the study of the surface properties of plasma treated polymers (Vargo and Gardella,
1989), in the investigation of the structure
of a self-assembled monolayer (King and
Czanderna, 1990) and in the determination
of failure locus in metal/polymer interfaces
(Puydtetal., 1988).
One should be aware that organic materials are generally sensitive to ion beam
damage and that sputtering is an unavoidable aspect of the ion beam-solid interaction at ISS beam energies (Gardella and
Pireaux, 1990). To avoid extensive damage, "static" or "low damage" conditions

(a)

2kV 3 H e + 1.45min 6.0nA/cm 2

600

700

800
900 1000 1100
Kinetic energy (eV)

1200

0.30

0.35

0.40

0.45
E/E o

0.50

0.55

0.60

Figure 21-26. Ion scattering spectra from poly(vinyl


pyridine), (a) P2VP, (b) P4VP, showing the sensitivity
of ion scattering spectroscopy (ISS) to the surface
structure of polymers. Reprinted with permission
from Hook et al. (1987). Copyright American Chemical Society.

have to be maintained (Gardella and Hercules, 1981). However, a recent study by


Puydt etal. (1990) of the extent of ion
bombardment damage to poly(methyl
methacrylate) and poly(ethylene terephthalate) showed that no "static" conditions for ISS analysis on polymers could be
achieved with acceptable signal-to-noise
ratios. The aspect of beam damage has
limited the use of ISS for surface selective
analysis of organic materials, particularly
in view of the fact that alternative methods
exist such as SSIMS, which exhibits high
elemental and chemical sensitivities for
surfaces. However, it should be noted that
ISS can be much more surface selective
than SSIMS and is directly sensitive to

734

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

structural variations of polymer surfaces


as shown by Gardella (1988). In principle,
one advantage of ISS should be the ability
to quantify surface composition. However,
this has been very difficult due to the uncertainties in absolute scattering cross sections and the lack of knowledge of the
probability of neutralization of the surface
scattered ions (Feldman and Mayer, 1986).
The poor signal-to-noise ratio as discussed
above adds further difficulty to the quantification of ISS data. It is, therefore not
expected that ISS will find major applications in the near future for the surface
analysis of organic materials relative to the
use of other techniques such as XPS and
SSIMS. However, ISS does have unique
advantages for surface structure and thereby will continue to be an occasional complementary tool to XPS and SIMS.
21.2.5.5 Secondary Ion Mass
Spectrometry (SIMS), Static SIMS
(SSIMS), Fast Atom Bombardment
Spectrometry (FABS)
and Dynamic SIMS (DSIMS)
The secondary ion mass spectrometry
techniques have received increasing attention as a major analysis method in the
study of the surfaces and interfaces of organic materials in recent years. For a detailed discussion of the uses of SIMS for
polymer analysis, the reader can refer to
several recent review papers (Briggs,
1989 a; van der Wei et al., 1990).
In a SIMS analysis, the sample is bombarded with a beam of neutral or charged
atoms at keV energies. At these energies,
the fundamental mechanism of interaction
of the probe particles with the solid is to
transfer large amounts of momentum into
the outer surface layers of the target solid
with the result that ejection of a mixture of
charged and neutral fragments from the

surface occurs (Smith etal., 1990; Foley


etal., 1984). The mechanism is not very
dependent upon the charge state of the
probe particle and consequently the fragment pattern is similar whether neutrals or
ions are used for the probes. In the simplest experiment the ejected ions (secondary ions) are mass-analyzed using one
of the three types of ion analyzers: time-offlight, magnetic sector and quadrapole.
Since the neutrals cannot be analyzed by
the ion detectors the total amount of information available in the typical experiment
is limited. However, in a more sophisticated experiment, the ejected neutral fragments also can be analyzed by ionizing
them immediately upon ejection from the
surface using various methods, typically
either electron or laser beams (Fulgham
etal., 1989; Winograd, 1993). The major
effect of this post-ionization is to increase
the total ion yield substantially, and thus
the signal-to-noise ratio, for a given probe
current. The SIMS experiment is called
static or dynamic according to the rate at
which the material is removed during analysis. In static SIMS (SSIMS), the numbers
of ions incident on the surface is very low
(on the order of < 10" 1 3 ions/cm2) and the
secondary ions come from the first one or
two layers of the material which are minimally damaged by the analysis. In dynamic
SIMS (DSIMS), the probe ion current is
much higher than in SSIMS, the surface is
rapidly eroded away and the analysis thus
is not restricted to the first layers. The material removal rate in SSIMS is typically
about 0.1 nm/h, while in DSIMS the surface is removed at the rate of about 10 |im/
h (Gabassi and Occhiello, 1987; Woodruff
and Delchar, 1986). The static mode is
highly preferred for careful analysis of the
chemical composition of the outer surface
since low ion doses ensure that the area to
be analyzed by an incoming ion will be free

21.2 Survey of Characterization Techniques

of damage from a previous bombardment


event. The dynamic mode has been applied
primarily to depth profile analyses of elemental composition. Molecular structure
and chemical bonding information is difficult to obtain in DSIMS because of the
severe damage which occurs during the
continual sputtering of the surface. SSIMS
is capable of providing both elemental and
molecular structure information of the
surface under investigation with a sensitivity of 1 0 " 5 - 1 0 " 6 atomic fractional coverage, a level unparalleled among surface
analysis techniques such as XPS, Auger
and ISS (Hoffmann, 1986). SSIMS is also
extremely surface selective with an analysis
depth of only about 1 nm (Hearn et al.,
1987). With the recent development of liquid metal ion sources capable of lateral
beam dimensions of 50 nm, high resolution imaging SIMS has become possible
(Winograd, 1993). However, since the ion
counts will be extremely low from such
small surface areas, extremely sensitive detectors are required to ensure that static
conditions are maintained. This problem
has been solved with the recent development of time-of-flight (TOF) detectors and
it is now possible to obtain good chemical
images of the surfaces of organic materials
(Eccles and Vickerman, 1989; van Ooij
et al., 1991; Hearn and Briggs, 1991; Briggs
etal., 1990).
There are, however, several drawbacks
associated with SSIMS. First, it is difficult
to quantify the atomic or molecular species
present on the surface from SSIMS measurements. This difficulty arises from
strong, highly specific matrix effects on the
secondary ion yield and the uncertainty in
the determination of the ionization probability which regulates the ratio of ions to
neutrals in the ejected fragments (Hearn
and Briggs, 1988; Feldman and Mayer,
1986). Second, the use of SSIMS on insu-

735

lating materials, which includes most organics, gives rise to surface charging. The
latter causes problems ranging from signal
instability to complete loss of signal in a
period of a few seconds (Bletsos et al.,
1988). Even in the case of TOF-SIMS,
where the primary ion currents are typically 1000 times less than those in quadrupole
SIMS, charging of insulators is still a problem (Eccles and Vickerman, 1989; Briggs
et al., 1990; Bletsos et al., 1988; Lub et al.,
1988). The charging problem can be partially or completely avoided by using neutral atoms instead of ions as the primary
beam for insulating materials (Bletsos
etal., 1988; Lub etal., 1988). The latter
technique has been referred to as fast atom
bombardment spectrometry (FABS) in the
literature. An additional advantage of using neutrals as the primary beam is the
increased secondary ion yield which enables static SIMS spectra to be obtained at
much reduced primary doses (van Ooij
etal., 1991). Surface charging also can be
compensated by supplying electrons to the
charged surface using a low energy electron flood gun. However, improper use of
a flood gun may give rise to electron stimulated ion emission (Briggs, 1986). Another difficulty associated with SSIMS is the
lack of understanding the relationship between the structure of the ionic fragments
obtained from the measurement and the
structure of the original surface. This
problem in particular makes it very difficult to obtain structure information from a
SSIMS analysis without extensive calibrations of the distinctive fingerprint mass
spectra from each specific type of material
(van der Wei et al., 1990).
Because of the rich detail of the spectra
obtained from organic materials, SSIMS
has been used extensively to study a variety
of problems associated with polymers. A
variety of pure polymers have been report-

736

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

ed in the literature and it is clear from these


studies that each polymer produces a
unique, fingerprint spectrum. The numbers of publications are too numerous to
list here. Instead, a few citations will be
made in order to illustrate the varieties of
polymers studied. The polymers investigated using SSIMS include polycarbonate
(Lub and Buning, 1990), acrylic homopolymers (Castner and Ratner, 1990),
polyurethane (Hearn etal., 1987), polystyrene (Lub et al., 1989), Nylon (Briggs,
1987), polyester (Briggs, 1986), and polyimide (van Ooij and Michael, 1990).
SSIMS has been used extensively in the
study of the surface-modified polymers
(Lub etal., 1989; Niehuis etal., 1989;
Brinen etal., 1991; Fowler etal., 1991).
Applications also have been found in the
investigation of metal-polymer interfaces
(van Ooij et al., 1991; Furman et al., 1990).
Although SSIMS has been used primarily as a qualitative technique to study the
surface chemistry of polymeric materials,
some quantitative applications exist. For
example, it has been shown that the composition of the components in random copolymers can be determined (Lub and
Buning, 1990; Castner and Ratner, 1990).
For the several copolymers investigated in
the latter studies, each monomer component appeared to generate a unique series
of fragment ions. The variations of the relative intensities of each of these characteristic monomer contributions to the total
copolymer spectra were observed to be
consistent with the compositional changes
in the copolymers. Copolymers investigated include n-butyl methacrylate-^-butyl
methacrylate copolymer (Briggs et al.,
1984), poly(styrene-/7-hydroxystyrene) copolymer (Chilkoti et al., 1990), ethyl
methacrylate - hydroxyethyl methacrylate
copolymer (Briggs and Ratner, 1988), nylon-6/nylon-66 copolymer (Briggs, 1986,

1987). With respect to the latter study,


Fig. 21-27 shows the SSIMS spectra of
both nylon-6 and nylon-66. It can be seen
that nylon-6 has a characteristic ion peak
at m/z = U4[M+H\+ where M is the mass
of the repeat unit of nylon-6 that is virtually absent in the nylon-66 spectrum and
both spectra show a common fragment at
m/z = 55. Using the common fragment at
m/z = 55 for normalization purposes produces a quantitative correlation as shown
in Fig. 21-28. Careful examination of all the
copolymer spectra reveals one departure
from the trend of the homopolymer spectral superposition which characterizes the
Nylon6-Nylon66 copolymers, the intensity
of a peak at m/z = 213. The latter feature is
very weak in the homopolymer spectra, but
reaches a significant intensity in the co-

209

50

100

150

200

210

200

Figure 21-27. Secondary ion mass spectroscopic


(SIMS) spectra of both (a) Nylon 6 and (b) Nylon 66.
Reprinted with permission from Briggs (1987). Copyright John Wiley & Sons Ltd.

21.2 Survey of Characterization Techniques

2.0 -

100
Nylon6 monomer (mol %)

Figure 21-28. Correlation of Secondary ion mass


spectroscopic (SIMS) relative peak intensities with
Nylon6-Nylon66 copolymer composition. Reprinted
with permission from Briggs (1987). Copyright John
Wiley & Sons Ltd.

polymers. This peak has been assigned as


+
CH 2 (CH 2 ) 4 NHCO(CH 2 ) 4 CONH 2 . The
peak intensity at m/z = 213, normalized using the intensity at m/z = 55, as a function
of nylon6-nylon66 copolymer composition is plotted in Fig. 21-28. Clearly a maximum can be seen in the I2\zl155 ratio. This
suggests that the ion responsible is derived
from a structural unit involving a link between the two monomers (Briggs, 1987).
While SSIMS has found wide applications in polymer surface analysis, dynamic
SIMS has only been used in a few cases.
The use has been limited largely because of
the complications involved with the ion
sputtering process of polymeric materials,
such as radiation damage accumulation,
preferential sputtering, surface morphology development and knock-in effects
(Lam, 1988; Liau et al., 1979). Depth reso-

737

lution in DSIMS is also significantly degraded as a function of combinations of


the above effects. The depth resolution has
been reported to be about 13 nm under
optimal condition (Russell, 1991; Coulon
et al., 1989). Like SSIMS, DSIMS also suffers from surface charging problems associated with lack of grounding of the ion
beam current through non-conducting
polymer materials. Deposition of a thin
layer of Au on top of the sample surface
has been used to achieve current leakage
and also to serve as a convenient marker
for the vacuum-polymer surface in
DSIMS experiments (Chou, 1990). Several
researchers have used DSIMS to investigate a variety of problems in polymer science such as the interdiffusion in homopolymers (Michael and Stuik, 1987; Whitlow and Wool, 1989), surface-induced ordering in symmetric diblock copolymers
(Coulon etal., 1989; Russell etal., 1989),
thickness measurement of lubricant films
on silicon and gold (Lorenz etal., 1991),
segregation of diblock copolymer into the
interface between two homopolymers (Deline etal., 1991) and the concentration
depth profile measurement of hydroxy-terminated polystyrene chains grafted unto a
native-oxide covered Si surface (Zhao
etal., 1991a). One of the important attributes of SIMS is the ability to analyze
H, a distinct advantage over other typical
surface techniques such as XPS which are
not sensitive to H atoms. As an example of
an application utilizing this advantage,
Coulon et al. (1989) have used DSIMS to
characterize the ordering in symmetric
diblock copolymers of polystyrene-poly(methyl methacrylate) in which one block
of the copolymer was deuterated. The concentration profile as a function of sputtering time for a deuterated PS-hydrogenated PMMA (50K/50K) diblock copolymer
is shown in Fig. 21-29. It can be seen from

738

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

10'

8
o

c
o

500

1000

1500

2000

Figure 21-30. Schematic representation of the surface-induced ordering process. Reprinted with permission from Coulon (1989). Copyright American
Chemical Society.

(sec)

Figure 21-29. Concentration profile as a function of


sputtering time for a deuterated polystyrene (PS)~
hydrogenated poly(methyl methacrylate) (PMMA)
diblock copolymer after annealing for 72 h at 170C.
Reprinted with permission from Coulon et al. (1989).
Copyright American Chemical Society.

the depth profile that order develops in the


form of periodicity of the microdomain
structures during annealing at 170 C. A
schematic of the surface-induced ordering
process in a diblock copolymer is shown in
Fig. 21-30 (Coulon etal., 1989). The top
diagram shows the diblock copolymer is in
a phase-mixed state. The center diagram
corresponds to a microphase-separated
morphology where the periodic lamellar
microdomains are randomly oriented in
the specimen. After annealing for 24 h at
170C, the copolymer exhibits a lamellar
morphology oriented to the surface as
shown in the bottom picture.
From the variety of studies reported in
the literature it can be concluded that

SIMS is a very sensitive technique for detection of elements and chemical groups in
organic surfaces. In the static mode the
technique exhibits high surface selectivity,
yielding information from the outer few
monolayers. Further, there are indications
that, with careful calibration, quantitative
measurements can be made. SIMS is a
highly complementary technique to XPS
since the SIMS spectrum contains completely different information than XPS, including information about the presence of
H and different functional groups. Thus in
cases for which chemical structures lead to
overlapping, unresolved core level peaks in
XPS, SIMS often can be used to distinguish between different possibilities, even
though the SIMS spectrum is very complex
and not fully understood. The complexity
of the spectra do represent a drawback to
the technique at present but it can be expected that as ongoing research continues
in the area of trying to understand the

21.2 Survey of Characterization Techniques

SIMS mechanism(s), it should be increasingly possible to elucidate the structures of


organic surfaces based on the fragments
produced and to obtain quantitative information on the numbers of molecules present in the outermost molecular layer of a
polymeric material (van der Wei et al.,
1990).
21.2.5.6 Atom Scattering
The scattering of atoms of thermal energies, particularly He atoms, from solid surfaces has developed into a very sensitive,
albeit specialized, technique for the analysis of the periodic structure of the very top
atomic layer of materials (Armand and
Salanon, 1989). The extreme surface selectivity of helium scattering arises from the
fact that for a low mass atom such as helium, striking a surface with the incoming
kinetic energies used in a typical scattering
analysis, there is no penetration of the
atom past the top atomic layer. At these
kinetic energies the He atom possesses a
wavelength suitable for diffraction from
atomic and molecular spacings so that the
use of a well-collimated beam allows diffractive scattering measurements from the
outermost surface layer. In this regard, helium diffraction is complementary to the
other surface diffraction probes such as
LEED and GIXRD, for which the probing
depth is of the order of a few nanometers.
The combination of all of these techniques
can provide a fairly complete evaluation of
the ordering in both the surface and interior of monolayers (Camillone et al., 1993 a).
Most importantly for analysis of organic
materials, the very low kinetic energy of
the atom beam used in helium scattering
analysis does not lead to damage of the
sample surface and the charge neutrality
precludes charge accumulation from occurring at the surface, a typical problem

739

with analysis by most surface techniques.


Such unique features have made this technique very useful in the study of the ordering of the outermost functional groups
of self-assembled monolayers (Camillone
etal., 1991, 1993ab; Chidsey et al., 1989)
and Langmuir-Blodgett films (Vogel and
Woll, 1988). Helium scattering has also
been used in the study of kinetics of adsorption of organic molecular species on
different substrates (Jung et al., 1991; Graham et al., 1992). A detailed discussion of
this technique can be found in a recent
book (Scoles, 1992). In general, because of
the requirement of smooth, planar, ordered organic surfaces for helium scattering analyses, the technique is restricted to
specialized samples such as LangmuirBlodgett films and crystalline self-assembled monolayers, as opposed to typical
polymer surfaces. Limited use for such
specialized samples is expected to continue.
21.2.6 Neutrons In, Neutrons Out:
Neutron Reflectivity (NR)
Neutrons are an interesting contrast to
other probe particles. Because their mass is
of the order of typical light elements (the
same as hydrogen), they carry similar momentum for given kinetic energies. But because neutrons contain no charges, their
collision interactions with solids depends
upon nuclear interactions since electrostatic forces can have no effects. Thus, the
collision mechanisms are completely different than for atoms or ions, which are
constituted of collections of charged particles. In particular, without electrostatic interactions, collision cross sections in a solid are very small and thus solids are far
more transparent to neutrons than to
atoms or ions (except at very high energies;
see Sec. 2.5). Further, damage is not gener-

740

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

ated along the trajectory through the solid


target, as occurs with atoms and ions, so
neutron probes are considered as nondestructive and, further, capable of insitu
analysis in gases and liquids. In some
sense, neutrons behave phenomenologically like photons, i.e., deep, nondestructive
penetration into solids. In addition, when
the quantum mechanical wave-like properties of the neutron are considered, the
analogy with light is very accurate. In particular, as with atoms, diffraction in solids
can occur and also solids can be assigned
effective refractive indices (real values see
Eq. (21-1)) which allow calculation of the
reflection properties with respect to neutron beams, exactly analogous to reflection properties of light (see Sec. 21.2.2).
Since neutrons typically have wavelengths
similar to X-rays, experiments quite
analogous to XRR (see Sec. 21.2.2) can be
carried out. This is the basis of NR and is
responsible for its ability to do nondestructive depth profiling and to characterize
interfaces in thin film structures. In particular, because of its nondestructive nature
and characteristic energy (wavelength)
range, NR is quite well suited to the study
of thin organic films and interfaces
(Kramer, 1991). However, in terms of the
analogy with light, the effective refractive
indices of the different constituents or layers of a sample must be different in order
to obtain reflection effects. For organic
materials and polymers, the most convenient contrast between two materials can
be provided by substitution of H by D,
which exhibit measurably different responses to neutrons. For samples with
suitable reflection properties, the technique can provide excellent depth resolution down to &1 nm with penetration
depths over hundreds of nanometers (Fernandez etal., 1988; Russell etal., 1988)
and with little damage to normally sensi-

tive polymeric materials (Russell, 1990).


The unprecedented depth resolution has
been used to study the initial stage of polymer interdiffusion, a challenge which no
other techniques have the capability to address (Stamm etal., 1991a), and to test
proposed theories of molecular motion,
such as the reptation model in polymers,
proposed originally by de Gennes (Felcher
etal., 1991). A detailed account of the
principles and the applications of NR in
the study of surfaces and interfaces in
polymers and molecular films is given in
recent review papers by Russell (1990) and
Penfold and Thomas (1990).
There are, however, some limitations associated with the NR technique. First, the
experiments can be done only in a few neutron reactor facilities located around the
world. Second, determination of density
profiles generally is not unique because
measurement of reflectivity by itself generates only one known and at least two
parameters are required for the calculation
(Russell, 1990). One solution to this problem is to couple the reflectivity measurements with other depth-profiling techniques, such as angle-resolved XPS,
DSIMS, FRES, NRA and ellipsometry
(Kramer, 1991; Russell, 1990; Jones et al.,
1990; Mansfield et al., 1991). Kramer has
discussed this aspect in detail (Kramer,
1991). Third, as with XRR, NR requires
extremely flat and smooth samples in order to be compatible with the needs for
accurate control of the reflection angles
and approach to grazing conditions. Thus,
specialized substrates such as extremely
flat glass and silicon with subnanometer
scale roughness is preferable which eliminates many samples of practical interest.
Finally, deuteration is often a necessary
step in sample preparation in order to
provide contrast for the reflectivity measurement.

21.2 Survey of Characterization Techniques

There have been an increasing number


of applications to thin organic films. The
first reports involved molecular Langmuir-Blodgett films (Nicklow et al., 1981;
Hayter etal., 1981) and these types of
studies have continued (Grundy et al.,
1988). The first reports of applications to
polymer films were more recent (Fernandez et al., 1988; Russell et al., 1988; Stamm
and Majkrzak, 1987). Several research
groups have made applications to the particularly complex problems of interdiffusion, interface structure and surface-induced ordering in polymer films. The
specific systems studied include: interdiffusion between the deuterated and hydrogenated components of the same polymers, such as d-PS-h-PS, d-polyimide-hpolyimide (Russell etal., 1988; Stamm
etal., 1991b; Felcher etal., 1991; Jones
et al., 1990); the interdiffusion between different polymers such as polystyrenepoly(methyl methacrylate), polystyrenepolybromostyrene
and
poly(methyl
methacrylate) -poly(vinyl chloride) (Fernandez etal., 1988; Zhao etal., 1991;
Stamm et al., 1991 b); the morphology and
ordering in block copolymers (Mansfield
et al. 1991; Anastasiadis et al., 1989,1990),
and the segregation of diblock copolymers
into the interface of the two homopolymers (Russell etal., 1991). Such applications are expected to continue and are
among the best uses of this technique.

21.2.7 Dynamic Depth Profiling


The analysis of a buried interface, viz,
one located between two adjoining material phases, is notoriously difficult due to the
difficulty in selectively analyzing the interface region. Ideally, the best sample geometry for interface analysis by surface selective techniques is a substrate covered with

741

an overlayer of material which is thinner


than the effective penetration depth of the
probe selected. For example, XPS analysis
(see Sec. 21.2.3.1) has been used to characterize the core levels of atoms in the region
of an interface which is located a few
atomic layers from the top surface and this
information then used to determine interfacial phenomena such as chemical reactions (Kowalczyk, 1990). General approaches to the problem of depth profiling, including optical methods, have been
reviewed (Bohn and Miller, 1991). However, in spite of the large number of very
useful surface selective analytical techniques, such as XPS, the interface selectivity constraint renders these techniques useless when the interface is buried below the
intrinsic probing depth.
A variety of "direct interface" techniques have been used by researchers to
obtain information about interfaces. These
essentially are techniques with low or no
surface selectivity, viz, large probing
depths, but with the ability to determine
composition or property gradients. These
techniques include XRR, SE, IRS, RBS,
FRES, NRA, and NR (see sections specifically dealing with these techniques). However, such techniques suffer certain disadvantages. First of all, with the exception of
IRS in the internal reflection mode
(Sec. 21.2.2.8), these techniques are not
sensitive to the chemical nature of the interface and IRS offers very poor depth resolution. Further, the optimum use of the
other techniques generally has not been
with organic materials.
An attractive alternate route to interface
analysis has been to artificially thin the
material to expose the interface and then
apply highly surface selective and chemically sensitive techniques such as XPS and
SSIMS (Watts, 1988; Castle and Watts,
1988). The most popular way of thinning

742

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

the material is by ion sputtering. Several


review papers have been published on various aspects of this technique (Coburn,
1976; Lam, 1988). Hashimoto (1991),
Vasile and Bachman (1989) have demonstrated that ion sputtering combined with
XPS can provide important chemical information about the buried interface between metals and polymers that cannot be
obtained otherwise.
However, caution should be exercised in
interpreting the experimental results when
sputtering is used for interface analyses
since permanent structural and chemical
changes can be introduced by the ion sputtering process. Those artifacts include: (1)
ion induced chemical reactions, i.e., the reduction of compounds from a higher oxidation state to a lower one and the formation of carbide when adventitious or constituent carbon is reacted with other components of the system under ion bombardment (Mitchill etal., 1990; Holm and
Storp, 1977; Selvam etal., 1990; Barr,
1977); (2) ion induced changes in composition, such as preferential sputtering and
the redistribution of elements due to the
"knock-on" effect (Holm and Storp, 1977;
Liau et al., 1979; Burrell et al., 1988; Contarini et al., 1987); and (3) ion induced microtopographical changes of the surface
(Kojima etal., 1987; Cirlin etal., 1990;
Nefedov, 1988).
Chemical changes at the interface
caused by ion bombardment has been
studied by several researchers. Holm and
Storp (1977), for example, have observed
that reactive metals can react with adventitious carbon during ion bombardment to
form metallic carbide. In the case of organic materials, similar destructive effects of
ion bombardment were observed. For an
organic material containing aromatic
rings, e.g., polystyrene, ion bombardment
causes the disappearance of the typical

shake-up satellite in the C Is range (Storp


and Holm, 1979).
Preferential sputtering and ion mixing
phenomena have been studied by several
research groups (Storp and Holm, 1979;
Williams and Davis, 1977; Cirlin etal.,
1990; Burrell etal., 1988; Contarini etal.,
1987). Preferential sputtering and ion mixing results in measured compositions
which deviate from the actual composition
of the system studied. Preferential sputtering of polymeric materials has been investigated (Nuzzo et al., 1987; Contarini
etal., 1987; Williams and Davis, 1977).
Rapid depletion of oxygen and nitrogen
has been observed in polyimide, poly(methyl methacrylate) and polyurethane
during the sputtering process and a carbon
rich surface has been observed after sputtering (Nuzzo et al., 1987; Williams and
Davis, 1977). It is generally accepted that
straightforward quantitative analysis of
most organic polymers is not feasible, even
though this approach is still useful to gain
a simple knowledge of the presence or absence of an element as a function of depth
(Williams and Davis, 1977).
The change of the surface morphology
due to the ion bombardment has a significant effect on the depth resolution of analysis and therefore results in artificial
broadening of the interface. Such broadening gives rise to increased difficulty in determining the true depth profile in interfacial analysis (Hofmann, 1992). The artificial broadening caused by ion bombardment can be reduced quite significantly by
making use of sample rotation of the sample (Cirlin et al., 1990).
Several workers have employed the ion
sputtering-surface analysis approach in
the analysis of metal-polymer interfaces.
Burrell etal. (1988, 1992), for example,
have studied the interface between copper
and poly(ester imide) films and have tried

21.3 Evaluation of Characterization Techniques

to correlate the interfacial chemistry with


the adhesion at metal-polymer interface.
Similar work has also been performed by
Andre et al. (1990), Hashimoto (1991),
and Vasile and Bachman (1989). A combination of ion sputtering and XPS has been
used in the analysis of the change of the
surface composition of a protective coating upon exposure to harsh environments
(Schwamm et al., 1991).
The use of ion sputtering automatically
generates fragments which can be analyzed
by mass spectrometry simultaneously. This
technique of dynamic secondary ion mass
spectrometry (DSIMS) has been used recently in the analysis of interfaces in organic materials such as polymers (see
Sec. 21.2.5.5). An example of a DSIMS
sputtering profile can be seen in Fig. 21-29.
Since a significant advantage of mass spectrometric detection is its very high sensitivity, trace element analysis can be performed in a reasonable time in a DSIMS
experiment. Another advantage of ion induced sputtering is the very high lateral
resolution ion beams offer in comparison
with other surface probes. This is a particularly important advantage for the analysis of interfaces in fiber-based composite
structures because of the small lateral dimension of fibers, generally tens of
micrometers. SIMS, however, suffers from
extreme difficulty in quantification of the
profiling data and the complexity of the
interpretation of the mass spectrum often
provides problems. However, it is generally believed that DSIMS has the potential
to provide significant information about
molecular structure profiles at organic material interfaces.

743

21.3 Evaluation of
Characterization Techniques
In the process of choosing and implementing a given analytical technique for
the solution of a particular surface-interface analysis problem, there are always
multiple criteria which must be considered.
While a specific technique usually can be
identified as the most ideal one for the
problem at hand, in practice, one inevitably finds that there are important factors which mitigate against the use of a
specific technique. The usual result is a
compromise selection of a technique which
is considered optimal, given all the various
factors playing in the decision. Seven factors, obvious to anyone who has worked in
the surface analysis field, which inevitably
become important considerations in the final decisions to pursue a given analysis
strategy are: (1) costs of purchase and operation of the instrument, if it is on site; (2)
travel convenience, if the instrument is not
located directly on site; (3) flexibility of
examining different shapes, sizes and types
of samples and the ease of sample preparation; (4) simplicity and ease of carrying out
the analysis measurement and the level of
expertise required; (5) level of information
content offered by the measurement relative to the level required to solve the problem; (6) level of quantitation offered relative to the level required to solve the problem; and (7) level of theoretical expertise
required to interpret the data in a useful
way. Certainly there will be other factors,
but the above ones are typically encountered across both industrial and academic
laboratories. The importance of each of
these factors will differ widely from user to
user and lab to lab but it is not unreasonable to try to assess an average importance
to each. With this in mind the authors,
based on their own experience and a

744

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

knowledge of experiences of many workers


in the field, have made a ranking of the
"raw" importance of each technique listed
in the previous sections with respect to the
ability of the technique, when applied in an
ideal circumstance, to provide useful and
detailed information about some aspect of
surfaces, interfaces and thin films of organic materials. This numerical ranking
then was modified by weighting each of the
above seven factors. This exercise was intended to "level up" the techniques' capabilities in terms of applications to a broad
range of realistic analysis strategies. The
final list is given in Table 21-4. Admittedly
the rankings are highly subjective but they
do end up reflecting quite well the general
day to day utility of a technique for organic surface-interface-thin film analysis in
a laboratory of modest means and expertise.

Table 21-4. Ranking of optimum techniques for the


characterization of surfaces, interfaces and thin films
of organic materials.
Ranking
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18

Technique (s)
XPS
IRS
SSIMS
VBXPS, SE, AFM
UPS, IETS
DSIMS
SFG
HREELS, SERS, STM
Raman, SPR, NEXAFS, RBS, FRES
AES
NRA
XRR, XRF
LEED
SHG, ISS
TED
GIXRD, EELS
UV-visible fluorescence
NR, He scattering

The top three tools, XPS, IRS and


SSIMS, are not surprising in view of their
proven utility, as judged from the literature and their common use in industrial
laboratories. In particular, all are sensitive
to the types of chemical structures present
in organic samples. However, some interesting tradeoffs should be noted. State-ofthe art XPS (monochromatized X-ray
source and high resolution analyzer) and
state-of-the art SSIMS (imaging TOFSIMS) instruments are expensive as laboratory items go, costing about an order of
magnitude more than a quality IRS instrument (Fourier transform, moderate wavelength resolution). However, as noted in
the previous sections, IRS, unlike the other
two techniques is not surface selective, other than at the jameter scale for the internal
reflection mode, and so only will yield surface information in the case of localized
surface layers of distinctly different composition from the bulk or for ultrathin film
samples. Thus the cost advantage of IRS is
offset by the loss in general ability to selectively gather surface information for a variety of sample types. Of the three instruments, TOF-SIMS is the most surface selective and perhaps the most sensitive to
chemical structure, but the complexity of
the mass spectra and the sputtering mechanisms, combined with the lack of ionization cross sections leave most of the information in the spectra unusable without extensive calibration work. All three instruments can be purchased in rather user
friendly configurations, software being a
dominant factor, and thus they are well
suited for a standard analysis laboratory.
Since these instruments are commonly
used, availability of trained personnel to
operate them and interpret data is reasonably good. IRS has the added advantage
that it can be used for a variety of nonsurface applications whereas XPS and SSIMS

21.3 Evaluation of Characterization Techniques

instruments are restricted to surface applications.


Right below the top three, the appearance of a number of tools may seem somewhat surprising to those familiar with surface analysis of inorganic materials, for
which other tools such as AES and ISS
might have been expected. The three very
different techniques, VBXPS, SE and
AFM appear together. VBXPS spectra can
be obtained with a good XPS instrument
and are more sensitive to chemical structure than XPS core level spectra. However,
the VB spectra are very complex and usually not very useful without expert analysis
and quantum chemical calculations, a system by system development. SE also requires extensive expertise to interpret the
data, but the data base and protocol necessary for interpretation is fairly developed,
in contrast to VBXPS, and highly quantitative analyses of smooth, planar thin film
structures are possible on a routine basis.
Further, SE can be used for a variety of in
situ studies of liquid interfaces. AFM,
while sensitive only to surface topography,
often minimal information for surface
chemistry problems, is on the other hand,
reasonably easy to use, applicable to many
sample types and relatively inexpensive,
costing in the range of a good IRS spectrometer. Next on the list is UPS and
IETS. UPS is similar to VBXPS while
IETS requires highly specialized fabrication of the sample but can provide very
rich vibrational spectra for a number of
practical films such as adhesives and coatings which can be fabricated into the IETS
sample. Continuing down the list, DSIMS
is one of the few techniques which can
provide compositional depth profiles and
see buried organic-organic interfaces.
SFG, although a very difficult experiment
to set up and perform is the only truly
interface selective experiment and provides

745

highly structurally sensitive vibrational


spectra, as does IETS and IRS. However,
it is applicable only to very limited types of
samples such as monolayers on smooth,
planar surfaces of certain materials, but
this is somewhat offset by the fact that it
can be used in the presence of liquids as
an in situ analysis. Likewise, HREELS,
SERS, STM, regular Raman spectroscopy
and SPR are applicable only to specialized
sample types, e.g., STM only works with
conducting samples and SERS requires
special morphology (rough or islanded),
coinage metal surfaces. NEXAFS requires
the use of a synchotron facility and so is
limited in convenience, although useful in
the same way as VBXPS and UPS. The
high energy ion scattering techniques of
RBS and FRS, when run in the most effective way, require a dedicated ion accelerator facility, which is perhaps 20 times the
cost of an IR. Thus in most cases one
would most likely not purchase an on-site
capability but rather would travel to an
established facility to carry out the analysis. A further set of tradeoffs involves the
offsetting factors: (1) the data are very easy
to interpret, require almost no calibrations
specific to a given sample set, provide accurate elemental compositions as a function of depth into the sample and can be
obtained from a variety of sample types,
including bulk polymers, and (2) no chemical information, except in rare cases, can
be obtained. Thus if a suitable facility can
be located it is very easy to obtain large
amounts of surface region elemental composition data as a standard measurement.
AES is an intermediate case. An Auger
capability is almost standard fare for a
UHV analysis chamber. Thus it is almost
always available in many analysis laboratories. However, although it is inexpensive
and easy to operate it usually does not give
much useful information for polymer and

746

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

molecular film analysis because of charging, electron beam-induced sample damage and the need for calibrations. While
the rest of the table could be discussed in
this manner, the discussion would be unduly long and at this point the reader can
discern the logic used. It is clear that the
rankings are very subjective. Realizing this
it should be understood that the intended
value of the rankings is to provide the
reader with a basis for attempting to make
decisions as to which tools should be considered to solve analysis problems when
the various types of analysis requirements
and resource constraints mentioned above
are operative.

21.4 References
Abrahan, F. R, Batra, I. P. S. (1989), Surf. Sci. 209,
L125.
Ahn, D. X, Franses, E. I. (1992), /. Phys. Chem. 96,
9952.
Akhter, S., Allan, K., Buchanan, D., Cook, J. A.,
Campion, A., White, J. M. (1988), Appl. Surf. Sci.,
55,241.
Allain, C , Aussere, D., Rondelez, F. (1981), Phys.
Rev. Lett. 49, 1694.
Allara, D. L. (1993), Crit. Rev. Surf Chem. 2, 91.
Allara, D. L., Nuzzo, R. G. (1985), Langmuir 1, 52.
Allara, D. L., Swalen, X D. (1982), /. Phys. Chem. 86,
2700.
Allara, D. L., White, C. W. (1978), in: Stabilization
and Degradation of Polymers, Advances in Chemistry Series 169: Allara, D. L., Hawkins, W. L. (Eds.)
Washington, DC: ACS.
Allara, D. L., White, D. W, Meek, R. L., Briggs, T.
H. (1976), /. Polym. Sci., Polym. Chem. Ed. 14, 93.
Allara, D. L., Murray, C. A., Bodoff, S. (1983), in:
Physicochemical Aspects of Polymer Surfaces,
Vol. 1, New York, Plenum, p. 33.
Allara, D. L., Teicher, D., Durana, X F. (1984), Chem.
Phys. Lett. 84, 20.
Amador, S. M., Pachence, J.M., Fischetti, R., McCauley, X P., Smith, A. B., Blasie, X K. (1993),
Langmuir 9, 812.
Anastasiadis, S. H., Russell, T. P., Satija, S. K.,
Majrzak, C. F. (1990), /. Chem. Phys. 92, 5677.
Anastasiadis, S. H., Russell, T. P., Satija, S. K.,
Majrzak, C. R (1989), Phys. Review Lett. 62, 1852.
Ancelein, H., Hauptman, Z. V., Banister, A. X,
Yarwood, X (1990), J. Polym. Sci, Polym. Phys. Ed.
28, 1611.

Anderson, R. X (1993), Crit. Rev. Surf. Chem. 2, 53.


Anderson, S. G., Leu, X, Silverman, B. D., Ho, P. S.
(1993), J. Vac. Sci. Technol. All, 368.
Andre, V., Arefi, R, Amouroux, X, Puydt, Y. D.,
Bertrand, P., Lorang, G., Delamar, M. (1989),
Thin Solid Films 181, 451.
Andre, V., Arefi, F , Amouroux, X, Lorang, G.
(1990), Surf. Interface Anal. 16, 241.
Apai, G., McKenna, W. P. (1991), Langmuir 7, 2266.
Armand, G., Salanon, B. (1989), Surf. Sci. 217, 317.
Aspnes, D. E. (1976), in: Optical Properties of Solids:
New Developments, Seraphin, B. O. (Ed.). Amsterdam: North-Holland, chapter 15.
Atanasoska, L., Cammarata, V., Stallman, B. X,
Kwan, W. S. V., Miller, L. L. (1992), Surf Interface
Anal. 18, 163.
Atanasoska, L., Anderson, S. G., Meyer, H. M.,
Weaver, X H. (1990), Vacuum 40, 85.
Atanasoska, L., Anderson, S. G., Meyer, H. M., Lin,
Z. D., Weaver, X H. (1987), J. Vac. Sci. Technol. A5,
3325.
Azzam, R. M. A., Bashara, N. M. (1977), Ellipsometry and Polarized Light, Amsterdam: North-Holland.
Baglin, X E. E., Williams, X S. (1989), in: Ion Beams
for Materials Analysis. Bird, X R., Williams, X S.
(Eds.). New York: Academic Press.
Bain, C. D., Whiteside, G. M. (1989), /. Phys. Chem.
93, 1670.
Bain, C. D., Davies, P. B., Ong, T. H., Ward, R. N.,
Brown, M. A. (1991), Langmuir 7, 1563.
Barr, T. L. (1977), J. Vac. Sci. Technol. 14, 660.
Barr, T. L. (1989), /. Vac. Sci. Technol. 7, 1677.
Barth, G., Linder, R., Bryson, C. (1988), Surf. Interface Anal. 11, 307.
Barton, S. W, Thomas, B. N., Flom, E. B., Rice, S.
A., Lin, B., Peng, X B., Ketterson, X B., Dutta, P.
(1988), /. Chem. Phys. 89, 2257.
Barton, S. W, Bosio, L., Cortes, R., Rondelez, R
(1992), Europhys. Lett. 17, 401.
Battistoni, C , Mattogno, G., Paparazzo, E. (1986),
Surf Interface Anal. 7, 117.
Batich, C. D. (1988), Appl. Surf. Sci. 32, 57.
Baun, W. L. (1982), Pure Appl. Chem. 54, 323.
Beamson, G., Briggs, D., Davies, S. R, Fletcher, I.
W, Clark, D. T., Howad, X, Gelius, U., Wannberg,
B., Balger, P. (1990), Surf Interface Anal. 15, 541.
Beamson, G., Bunn, A., Briggs, D. (1991), Surf. Interface Anal. 17, 105.
Bee, T. G., McCarthy, T. X (1992), Macromolecules
25, 2093.
Bellamy, L. X (1975), The Infrared Spectra of Complex Molecules. New York: X Wiley.
Belton, P., Saffa, A., Wilson, R. (1988), Proc. Int.
Conf Anal. Appl. Spectrosc: Creaser, C , Davies,
A. (Eds.). London: Royal Soc. Chem., p. 245.
Bent, B. E., Somorjai, G. A. (1989), Adv. Colloid
Interface Sci., 29, 223.
Berthou, H., Xorgensen, C. K. (1975). Anal. Chem. 37,
482.

21.4 References

Bhatia, Q. S., Pan, D. H., Koberstein, J. T. (1988),


Macromolecules 21, 2166.
Binnig, G., Rohrer, H., Gerber, Ch., Weibel, E.
(1982), Phys. Rev. Lett. 49, 57.
Binnig, G., Quate, C. F., Gerber, Ch. (1986), Phys.
Rev. Lett. 56, 930.
Bird, J. R., Willams, J. S. (Eds.) (1989), Ion Beams for
Materials Analysis. New York: Academic Press.
Bletos, I. V., Hercules, D. M., Magill, J. H., van
Leyen, D., Niehuis, E. N., Benninghoven, A.
(1988), Anal. Chem. 60, 938.
Bloch, J. M., Sansone, M., Rondelez, R, Peiffer, D.
G., Pincus, P., Kim, M. W, Eisenberger, P. M.
(1985), Phys. Rev. Lett. 54, 1039.
Boerio, F. X, Hong, P. P., Tsai, H. W., Young, J. T.
(1991), Surf. Interface Anal. 17, 448.
Boerma, D. O. (1990), NucL Instrum. Methods Phys.
Res. B50, 11.
Bohn, P. W. (1988), Spectroscopy 3, 38.
Bohn, P. W, Miller, D. R. (1991), Crit. Rev. Anal.
Chem. 22, 1.
Bohn, P. W, Walls, D. J. (1991), Mikrochim. Acta 1, 3.
Born, M., Wolf, E. (1975), Principles of Optics, 5th
ed. New York: Pergamon Press.
Bou, M., Martin, J. M., Mogne, T. L., Vovelle, L.
(1991), Appl. Surf Sci. 47, 149.
Boulanger, P., Riga, J., Verbist, J. J., Delhalle, J.
(1989), Macromolecules 22, 173.
Boulanger, P., Magermans, C , Verbist, J. J., Delhalle,
I, Urch, D. S. (1991), Macromolecules 24, 2757.
Briber, R. M., Khoury, F. (1988), J. Polym. Sci.
Polym. Phys. (Ed.) 26, 621.
Briggs, D. (1986), Surf Interface Anal. 9, 391.
Briggs, D. (1987), Mass Spectrom 22, 91.
Briggs, D. (1989), in: Encyclopedia of Polymer Materials Science. New York: Wiley.
Briggs, D. (1989a), British Polym. J. 21, 3.
Briggs, D., Hearn, M. J. (1985), Spectrochim. Acta
40B, 707.
Briggs, D., Ratner, B. D. (1988), Polym. Commun. 29,
6.
Briggs, D., Seah, M. P. (Eds.) (1990), Practical Surface Analysis by Auger and X-ray Photoelectron
Spectroscopy. New York: Wiley.
Briggs, D., Hearn, M. X, Ratner, B. D. (1984), Surf.
Interface Anal. 6, 184.
Briggs, D., Hearn, M. X, Fletcher, I. W, Waugh, A.
R., Mclntoch, B. X (1990), Surf Interface Anal. 15,
62.
Brinen, X S., Greenhouse, S., Pinatti, L. (1991), Surf
Interface Anal. 17, 63.
Brito, A. N., Keane, M. P., Correia, N., Svensson, S.,
Gelius, U., Lindberg, B. X (1991), Surf Interface
Anal, 17, 94.
Browning, R. (1987), MRS Bulletin 7, 75.
Bryant, M. A., Pemberton, X E. (1991), J. Am. Chem.
Soc. 113, 3629.
Bryant, M. A., Pemberton, X E. (1991a), J. Am.
Chem. Soc. 113, 8284.

747

Buchwalter, L. P., Czornyi, G. (1990), /. Vac. Sci.


TechnoL A8, 781.
Buckley, C. P., Taylor, R. X (1984), J. Appl. Polym.
Sci. 29, 1985.
Buhaenko, M. R. (1988), Thin Solid Films, 159, 253.
Bukowska, X, Kudelski, A., Jackowska, K. (1991), J.
Electroanal. Chem. 309, 251.
Burkstrand, X M. (1981), J. Appl. Phys. 52, 4795.
Burrell, M. C , Porta, G. M., Karas, B. R., Foust, D.
F., Chera, X X (1992), J. Vac. Sci. TechnoL AW,
2752.
Burrell, M. C , Fontana, X, Chera, X X V. (1988), /.
Vac. Sci. TechnoL A6, 2893.
Chain, S. R. (1988), Chem. Phys. Lett. 143, 361.
Camillone, N. Ill, Chidsey, C. E. D., Li, G. Y,
Putvinski, T. M., Scoles, G. (1991), J. Chem. Phys.
94, 8493.
Camillone, N. Ill, Chidsey, C. E. D., Eisenberger, P.,
Fenter, P., Li, X, Liang, K. S., Liu, G. Y, Scoles, G.
(1993 a), J. Chem. Phys. 99, 744.
Camillone, N. Ill, Chidsey, C. E. D., Liu, G. Y,
Scoles, G. (1993 b), J. Chem. Phys. 98, 3503.
Campbell, D. X, Higgins, D. A., Corn, R. M. (1990),
/. Phys. Chem. 94, 3681.
Cartier, E., Pfluger, P., Pireaux, X X, Vilar, M. R.
(1987), Appl. Phys. A44, 43.
Castle, X E., Watts, X F. (1988), in: Interfaces in Polymer, Ceramic and Metal Matrix Composites, Elsevier, New York.
Castner, D. G., Ratner, B. D. (1990), Surf Interface
Anal. 15, 479.
Cazaux, X, Collieux, C. (1990), J. Electron Spectrosc.
Relat. Phenom. 52, 837.
Cazaux, X, Lehuede, P. (1992), J. Electron Spectrosc.
Relat. Phenom. 59, 49.
Cederberg, A. A. (1981), Surf Sci. 103, 148.
Chabal, Y (1988), Surf. Sci. Rep. 8,211.

Chang, C. C. (1987), MRS Bulletin 7, 70.


Charmet, X C , de Gennes, P. G. (1983), /. Opt. Soc.
Am. 73, 1777.
Chaturvedi, U. K., Steiner, U., Zak, O., Krausch, G.,
Klein, X (1989), Phys. Rev. Lett. 63, 616.
Chaturvedi, U. K., Steiner, U., Zak, O., Krausch, G.,
Shatz, G., Klein, X (1990), Appl. Phys. Lett. 56,
1228.
Chauvin, C , Sacher, E., Yelon, Y (1987), in: Surface
and Colloid Science in Computer Technology. Mittal, K. L. (Ed.). New York: Plenum.
Chemla, D. S., Zyss, X (1987), Nonlinear Optical
Properties of Organic Molecules and Crystals. New
York: Academic.
Chen, S. H., Frank, C. W (1989), Langmuir 5, 978.
Chiang, S., Tobin, R. G., Richards, P. L., Thiel, P. A.
(1987), Phys. Rev. Lett. 52, 648.
Chidsey, C. E. D., Liu, G. Y, Rowntree, P. A., Scoles,
G. (1989), /. Chem. Phys. 91, 4421.
Chilkoti, A., Ratner, B. D. (1991), Surf. Interface
Anal. 17, 567.
Chilkoti, A., Castner, D. G., Ratner, B. D., Briggs,
D. (1990), J. Vac. Sci. TechnoL A8, 2274.

748

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

Chtaib, M., Ghijsen, X, Pireaus, J. X, Caudano, R.,


Johnson, R. L., Orti, E., Bredas, X L. (1991), Phys.
Rev. B. 44, 10815.
Chou, N. X (1990), in: New Characterization Techniques for Thin Polymer Films'. Tong, H. M.,
Nguyen, L. T. (Eds.). New York: Wiley.
Chu, W. K., Mayer, X W., Nicolet, M. A. (1978),
Backscattering Spectrometry. New York: Academic Press.
Cirlin, E. H., Cheng, Y. T., Ireland, P., Clemens, B.
(1990), Surf. Interface Anal. 15, 142.
Clark, D. K., Dilks, A. (1977), /. Polym. Sci. Polym.
Chem. Ed. 15, 15.
Clark, D. T., Harrison, A. (1981), /. Polym. Sci.
Polym. Chem. Ed. 19, 1945.
Clark, D. T, Fok, Y. C. T., Roberts, G. G. (1981), /.
Electron Spectrosc. Relat. Phenom. 22, 173.
Coburn, X W. (1976), J. Vac. Sci. Technol. 13, 10371044.
Colletti, R. K, Gold, H. S., Dybowski, C. (1987),
Appl. Spectroscopy 41, 185.
Collins, R. W. (1990), Rev. Sci. Instrum. 61, 2029.
Collins, R. W, Allara, D. L., Kim, Y. T., Lu, Y, Shi,
X (1993), in: Characterization of Ultrathin Organic
Films'. Ulman, A. (Ed.) Greenwich, CT: Manning,
in press.
Composto, R. X, Kramer, E. X, White, D. M. (1990),
Polym. 31, 2320.
Compton, S. V., Compton, D. A. C , Messerschmidt,
R. G. (1991), Spectroscopy, 35.
Comyn, X, Horley, C. C , Oxley, D. P., Pritchard, R.
G., Tegg, T. L. (1983), The Brit. Polym. J. 15, 50.
Comyn, X, Horley, C. C , Oxley, D. P., Pritchard, R.
G., Tegg, T. X (1981), / Adhesion 12, 111.
Contarini, S., Schultz, X A., Tacchi, S., Yo, Y S.,
Rabalais, X W. (1987), Appl. Surf Sci. 28, 291.
Corn, R. (1991), Anal. Chem. 63, 285 a.
Cota, L., Adem, E., Yacaman, M. X (1986), Appl.
Surf. Sci. 27, 106.
Coulon, G., Russell, T. P., Deline, V. R. (1989),
Macromolecules 22, 2581.
Coxon, P., Krizek, X, Humpherson, M., Wordell, I.
R. M. (1990), J. Electron Spectrosc. Relat. Phenom.
52, 821.
Cros, A. (1992), J. Electron Spectrosc. Relat. Phenom.
59, 1.
Culler, S. R., Ishida, H., Koenig, X L. (1983), Ann.
Rev. Mater. Sci. 13, 303.
Currie, X, Depelsenaire, P., Groleau, R., Sacher, E.
(1984), /. Colloid Interface Sci. 97, 410.
Da Silva, L. B., Trebes, X E., Balhorn, R., Mwrowka,
S. (1992), Science 258, 269.
Davenas, X, Xu, X. L., Martrot, M., Mathis, C ,
Francois, B. (1988), Nucl. Instrum. Phys. Res. B32,
166.
Davenas, X, Xu, X. L., Boiteux, G., Sage, D. (1989),
Nucl. Instrum. Phys. Res. B39, 754.
Debe, M. (1987), Prog. Surf. Sci. 24, 1.
Debe, M. K., Field, D. R. (1991), J. Vac. Sci. Technol.,
A9, 1265.

deBruijn, H. E., Altenburg, B. S. F., Kooyman, R. P.


H., Greve, X (1991), Optics Commun. 82, 425.
Delhalle, X, Delhalle, S., Riga, X (1987), /. Chem. Soc.
Faraday Trans. 2 83, 503.
Deline, V. R., Brown, H. R., Char, K. (1991), J. Vac.
Sci. Technol. A9, 1283.
Dierker, S. B., Murray, C. A., Legrange, X D., Schlotter, N. E. (1987), Chem. Phys. Lett. 137, 453.
Dilks, A. (1981), in: Photon, Electron and Ion Probes
of Polymer Structure and Properties, ACS Sympos.
Series 162: Dwight, D. W, Fabish, T. X, Thomas,
H. R., Washington, DC: ACS.
Dorset, D. L. (1990), Macromolecules 23, 623.
Dorset, D. L. (1991), Ultramicroscopy 38, 23.
Doyle, B. L., Walsh, D. S., Lee, S. R. (1991), Nucl.
Instrum. Meth. Phys. Res. B54, 244.
Drevillon, B. (1988), Thin Solid Films 163, 157.
Dubois, L. (1993), in: Characterization of Organic
Thin Films, Ulman, A. (Ed.). Greenwich, C. T.:
Manning, in press.
Dubois, L. H., Nuzzo, R. G. (1992), Annu. Rev. Phys.
Chem. 43, 437.
Dubois, L. H., Zegarski, B. R., Nuzzo, R. G. (1993),
/. Chem. Phys. 98, 678.
Duke, C. B., Salaneck, W. R., Fabish, T. X, Ritsko, X
X, Thomas, H. R., Paton, A., Phys. Rev. B 18,
5717.
Dwight, D. W, McGrath, X E., Riffle, X S., Smith, S.
D., York, G. A. (1990), /. Electron Spectrosc. Relat.
Phenom. 52, 457.
Ebel, H., Ebel, M. F , Krocza, H. (1988), Surf. Interface Anal. 12, 137.
Eccles, A. X, Yickerman, X C. (1989), J. Vac. Sci.
Technol. A7, 234.
Eisenthal, K. B. (1992), Annu. Rev. Phys. Chem. 43,
627.
Endisch, D., Rauch, F , Gotzleman, A., Reith, G.,
Stamm, M. (1992), Nucl. Instrum. Methods Phys.
Res. B62, 513.
Erskine, X L. (1986), /. Vac. Sci. Technol. 4A, 1282.
Factor, B. X, Russell, T. P., Toney, M. F. (1991), Phys.
Rev. Lett. 66, 1181.
Fadley, C. S. (1984), Prog. Surf. Sci. 16, 275.
Fauster, T. (1988), Vacuum 38, 129.
Feidenhans'l, R. (1989), Surf. Sci. Rep. 10, 105.
Felcher, G. P., Karim, A., Russell, T. P. (1991), J.
Non-Cryst. Solids 131/133, 703.
Feldman, L. C , Mayer, X W (1986), Fundamentals of
Surface and Thin Film Analysis. Amsterdam:
North-Holland.
Fenter, P., Eisenberger, P., Li, X (1991), Langmuir 7,
2013.
Fenter, P., Eisenberger, P., Liang, K. S. (1993), Phys.
Rev. Lett. 70, 2447.
Fernandez, M. L., Higgins, X S., Penfold, X, Ward, R.
C , Shockleton, C , Walsh, D. X (1988), Polymer 29,
1923.
Ferrieu, F., Dutarte, D. (1990), J. Appl. Phys. 68,
5810.

21.4 References

Firment, L. E., Somorjai, G. A. (1979), Isr. J. Chem.


18, 285.
Roley, K. E., Winograd, N., Garrison, B. X, Harrison, D. E. (1984), /. Chem. Phys. 80, 5254.
Fowler, D. E., Johnson, R. D., van Leyen, D., Benninghoven, A. (1991), Surf. Interface Anal. 17, 125.
Foster, M., Stamm, M., Reiter, G., Huttenbach, S.
(1990), Vacuum 41, 1441.
Frank, L. (1991), Vacuum 42, 147.
Frantz, P., Granick, S. (1991), Phys. Rev. Lett. 66,
899.
Friend, C. M. (1989), in: Adhesion and Friction:
Grunze, M., Kreuzer, H. J. (Eds.). Berlin: SpringerVerlag.
Frommer, J, (1992), Angew. Chem. Int. Ed. Engl. 31,
1298.
Fulgham, J. E., McGuire, G. E., Musselman, I. H.,
Nemanich, R. X, White, X M., Chopra, D. R.,
Chourasia, A. R. (1989), Anal. Chem. 61, 243R.
Furman, B. K., Purushothaman, S., Castellani, E.,
Renick, S., Neugroshl, D. (1990), in: Metallization
of Polymers, ACS Symp. Series 440. Washington,
DC: ACS.
Gabassi, K, Occhiello, E. (1987), Anal. Chim. Acta
197, 1.
Gall, T. P., Kramer, E. X (1991), Polym. 32, 2320.
Gardella Jr., X A. (1988), Appl. Surf. Set 31, 72.
Gardella Jr., X A., Hercules, D. M. (1981), Anal.
Chem. 53, 1879.
Gardella Jr., X A., Pireaux, X X (1990), Anal. Chem.
62, 645A.
Gardella Jr., X A., Chin, R. L., Ferguson, S. A.,
Farrow, M. M. (1984), /. Electron Spectrosc. Relat.
Phenom. 34, 97.
Gardella Jr., X A., Ferguson, S. A., Chin, R. L.
(1986), Appl. Spectrosc. 40, 224.
Garoff, S., Deckman, H. W (1986), /. Phys. Paris 47,
701.
Gelius, U., Heden, P. F , Hedman, X, Lindberg, B. X,
Manne, R., Nordberg, C , Siegbahn, K. (1970),
Phys. Scr. 2, 70.
Gelius, U., Wannberg, B., Baltzer, P., Feldegg, H. F ,
Carlsson, G., Johansson, C. G., Larsson, X,
Munger, P., Vegefors, G. (1990), /. Electron Spectrosc. Relat. Phenom. 52, 141
Gerenser, L. X (1988), J. Vac. Sci. Technol. A6, 2897.
Goldberg, M. X, Clabes, X G., Kovac, C. A. (1988),
/. Vac. Sci. Technol. A6, 991.
Gossmann, H. X, Feldman, L. C. (1987), MRS Bulletin 8, 26.
Graf, R. T.. Eng, R, Koznig, X L., Ishida, H. (1986),
Appl. Spectrosc. 40, 498.
Graham, A. P., Allison, W, McCash, E. M. (1992),
Surf Sci. 269/270, 394.
Grant, X T. (1989), Surf. Interface Anal. 14, 211.
Green, P. F , Berger, L. L. (1993), Thin Solid Films
224, 209.
Green, P. F., Doyle, B. L. (1990), in: New Characterization Techniques for Thin Polymer Films, Tong,
H. M., Nguyen, G. T. (Eds.). New York: Wiley.

749

Green, P. R, Russell, T. P. (1991), Macromolecules


24, 2931.
Griffiths, P. R., DeHaseth, X A. (1986), Fourier
Transform Infrared Spectroscopy, Chemical Analysis, Vol. 83. New York: Wiley.
Groleau, R., Gurjrathi, S. C , Martin, X (1983), Nucl.
Instrum. Methods Phys. Res. 218, 11.
Grundy, M. X, Richardson, R. M., Roser, S. X, Penfold, X, Ward, R. C. (1988), Thin Solid Films 159,
43.
Grunthaner, F. X (1987), MRS Bulletin 8/9, 60.
Guntherodt, H. X, Wiessendanger, R. (Eds.) (1992),
Scanning Tunneling Microscopy. Berlin: SpringerVerlag.
Gui, X Y, Stern, D. A., Frank, D. G., Lu, R, Zapien,
D. C , Hubbard, A. T. (1991), Langmuir 7, 955.
Gujrathi, S. C. (1990), in: Metallization of Polymers,
ACS Symp. Series 440. Washington, DC: ACS.
Guyot-Sionnest, P., Hunt, X H., Shen, Y. R. (1987),
Phys. Rev. Lett. 59, 1597.
Guzonas, D. A., Hair, M. L., Tripp, C. P. (1990),
Appl. Spectrosc. 44, 290.
Hansma, P. K. (1987), in: Vibrational Spectroscopy of
Molecules on Surfaces. New York: Plenum.
Hansma, P. K. (1977), Phys. Rep. 30, 145.
Harrick, N. X (1967), Internal Reflection Spectroscopy, New York: Wiley.
Harris, A. L., Chidsey, C. E. D., Levinos, N. X, Loiacono, D. N. (1987), Chem. Phys. Lett. 141, 350.
Harris, A. L., Rothberg, L., Dhar, L., Levinos, N. X,
Dubois, L. H. (1991), /. Chem. Phys. 94, 2438.
Hashimoto, S. (1991), Appl. Surf. Sci. 47, 323.
Hayter, X B., Highfield, B. R., Pullman, B. X, Thomas, R. K., McMullen, A. I., Penfold, X (1981), /.
Chem. Soc. Faraday Trans. I 77, 1437.
Hearn, M. X, Briggs, D. (1988), Surf. Interface Anal.
11, 198.
Hearn, M. X, Briggs, D. (1991), Surf Interface Anal.
17, 421.
Hearn, M. X, Briggs, D., Yoon, S. C , Ratner, B. D.
(1987), Surf. Interface Anal. 10, 384.
Hedman, X, Heden, P. F , Nordberg, R., Nordling,
C , Lindberg, B. X (1970), Spectrochim. Acta 26A,
761.
Henriksen, P. N., Gent, A. N., Ramsier, R. D., Alexander, X D. (1988), Surf. Interface Anal. 11, 283.
Herminghaus, S., Leiderer, P. (1989), Appl. Phys.
Lett. 54, 99.
Herminghaus, S., Leiderer, P. (1991), Appl. Phys.
Lett. 58, 352.
Ho, P. S., Hahn, P. O., Bartha, X W, Rubloff, G. W,
LeGoues, R K., Silverman, B. D. (1985), /. Vac.
Sci. Technol. A3, 739.
Hobbs, X, Sung, C. S. P., Krishnan, K., Hill, S.
(1983), Macromolecules 16, 183.
Hoffmann, S. (1986), Surf. Interface Anal. 9, 3.
Hofmann, S. (1992), /. Vac. Sci. Technol. B10, 316.
Holm, R., Storp, S. (1977), Appl. Phys. 17, 101.
Hong, P. P., Boerio, R X, Clarson, S. X, Smith, S. D.
(1991), Macromolecules 24, 4770.

750

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

Hook, T. X, Gardella Jr., J. A. (1987), /. Vac. Sci.


Technol. A5, 1332.
Hook, T. X, Gardella Jr., J. A., Salvati, L. S. (1987a),
Macromolecules 20, 2112.
Hook, T. X, Gardella Jr., J. A., Salvati Jr., L. S.
(1987 b), /. Mater. Res. 2, 132.
Hook, T. J., Schmitt, R. L., Gardella Jr., J. A., Salvati
Jr., L. S., Chin, R. L. (1986), Anal. Chem. 58, 1285
Huang, X., Burzynski, R., Prasad, P. N. (1989),
Langmuir 5, 325.
Ibach, H. (1977), Surf. Sci. 66, 56.
Ibach, H., Mills, D. L. (1982), Electron Energy Loss
Spectroscopy and Surface Vibration. New York:
Academic.
Ishida, H. (1987), Rubber Chem. Technol. 60, 497.
Johnson, H., Granick, S. (1992), Science 255, 966.
Jones, R. A. L., Kramer, E. X, Rafailovich, M. H.,
Sokolov, X, Schwarz, S. A. (1989), Phys. Rev. Lett.
62, 280.
Jones, T. S., Ashton, M. R., Richardson, N. V.,
Unertl, W. N. (1989 a), J. Phys. Condens. Matt. 1,
SB139.
Jones, R. A. L., Norton, L. X, Kramer, E. X, Composto, R. X, Stein, R. S., Russell, T. P., Mansour, A.,
Karim, A., Felcher, G. P., Rafailovich, M. H.,
Sokolov, X, Zhao, X., Schwarz, S. A. (1990), Europhys. Lett. 12, 41.
Jordan-Sweet, X L. (1990), in: Metallization of Polymers, ACS Symp. Series 440. Washington, DC:
ACS.
Jordan-Sweet, X L.., Kovac, C. A., Goldberg, M. X,
Morar, X F. (1988), /. Chem. Phys. 89, 2482.
Jordan-Sweet, X L., Sanda, P. N., Morar, X F., Kovac, C. A., Himpsel, F. X, Pollak, R. A. (1988), /.
Vac. Sci. Technol. A4, 1049.
Jung, D. R., Cui, X H., Frankl, D. R. (1991), Phys.
Rev. B 43, 10042.
Kawaguchi, M., Nagata, K. (1991), Langmuir, 7,
1478.
Kelley, M. X (1991), MRS Bulletin 13, 46.
Kelley, K., Ishino, Y, Ishida, H. (1987), Thin Solid
Films 154, 271.
Kim, M. W., Pfeiffer, D. G., Chen, W, Hsiung, H.,
Rasing, T, Shen, Y R. (1989), Macromolecules 22,
2682.
Kim, Y T., Collins, R. W, Vedam, K., Allara, D. L.
(1991), /. Electrochem. Soc. 138, 3266.
Kim, Y T, Bard, A. J. (1992), Langmuir 8, 1096.
Kimura, F , Umemura, T., Takenaka, T. (1986), Langmuir 2, 96.
King, D. E., Czanderna, A. W. (1990), Surf. Sci. Lett.
235, L329.
Klein, X (1991), / Non-Cryst. Solids 131/133, 598.
Knoll, W, Philpott, M. R., Golden, W. G. (1982), J.
Chem. Phys. 77, 219.
Koh, S. L., Pae, K. D., Stoffel, N. G., Hart, D. L.
(1990), Polym. Eng. Sci. 30, 137.
Kojima, I., Fukumoto, N., Kurahashi, M. (1987), /.
Electron Spectrosc. Relat. Phenom. 50, C9.

Kooyman, R. P. H., Krull, U. X (1991), Langmuir 7,


1506.
Korzeniewski, C , Pons, S. (1987), Prog. Anal. Spectrosc. 10, 1.
Kowalczyk, S. P. (1990), in: Metallization of Polymers, ACS Symp. Series 440, Washington, DC:
ACS.
Kowalczyk, S. P., Jordan-Sweet, X L. (1990), Chem.
Mater. 1, 592.
Kramer, E. X (1991), Physica B 173, 189.
Krieger, U. K., Lanford, W A. (1988), J. Non-Cryst.
Solids 102, 50.
Krishna, M. R., Kallury, M., Thompson, M. (1992),
Langmuir 8, 947.
Laibinis, P. E., Bain, C. D., Whiteside, G. M. (1991),
/. Phys. Chem. 95, 7017.
Laibinis, P. E., Whitesides, G. M., Allara, D. L., Tao,
Y T., Parikh, A. N., Nuzzo, R. G. (1991), /. Am.
Chem. Soc. 113, 7152.
Lam, N. Q. (1988), Surf Interface Anal. 12, 65.
Lambe, X, Jacklevic, R. C. (1968), Phys. Rev. 165,
821.
Lanford, W A. (1978), Nucl. Instrum. Methods 149,1.
Lanford, W A., Zeigler, X F , Keller, X (1976), Appl.
Phys. Lett. 20, 566.
Lasky, R. C , Kramer, E. X, Hui, C. Y (1988), Polymer 29, 1131.
Lazzaroni, R., Bredas, X L., Dannetun, P., Logdlung,
M., Uvdal, K., Salaneck, W. R. (1991), Synth.
Metals. 41/43, 3323.
Lenk, T. X, Hallmark, V. M., Rabolt, X F., Haussling,
L., Ringsdorf, H. (1993), Macromolecules 26,1230.
Leung, O. M., Goh, M. C. (1992), Science 255, 64.
Leyden, D., Murthy, R. (1988), Trends. Anal. Chem.
7, 164.
Liau, Z. L., Tsaur, B. Y, Mayer, J. W. (1979), /. Vac.
Sci. Technol. 16, 121.
Liau, Z. L., Tsaw, B. Y, Mayer, X W (1979), / Vac.
Sci. Technol. 16, 121.
Liehr, M., Thiry, P. A., Pireaux, X X, Candano, R.
(1986), Phys. Review B33, 5682.
Lin, S. S. (1987), Surface Interface Anal. 10, 110.
Lin-Vien, D., Colthup, N. E., Fateley, W G., Graselli, J. G. (1992), The Handbook of Infrared and
Raman Characteristic Frequencies of Organic
Molecules. New York: Academic Press.
Lindberg, B. X, Hamrin, K., Johansson, G., Gelius,
U., Fahkman, A., Nordling, C , Siegbahn, K.
(1970), Phys. Scr. 1, 286.
Lofas, S., Malmqvist, M., Ronnberg, I., Stenberg, E.,
Liedberg, B., Lundstrom, I. (1991), Sensors and
Actuators B5, 79.
Lopez, G. P., Castner, D. G., Ratner, B. D. (1991),
Surf. Interface Anal. 17, 267.
Lopez, G. P., Dwight, D. W, Polk, M. B. (1986),
Surf. Interface Anal. 9, 405.
Lorenz, M. R., Novotny, V. X, Deline, V. R. (1991),
Surf. Sci. 250, 112.
Lotz, B., Wittmann, X C , Stocker, W, Magnonov, S.
N., Cantow, H. J. (1991), Polym. Bull. 26, 209.

21.4 References

Lub, J., Buning, G. H. W (1990), Polym. 31, 1009.


Lub, X, Velzen, P. N. T., van Leyen, D., Hagenhoff,
B., Benninghoven, A. (1988), Surf. Interface Anal.
12, 53.
Lub, I, Vroohoven, F. C. B. M., Bruninx, E., Benninghoven, A. (1989), Polym. 30, 40.
Mack, R. G., Grossman, E., Unertl, W. N. (1990), J.
Vac. ScL Technol. A8, 3827.
Mackenzie, M. W. (Ed.) (1988), Advances in Applied
Fourier Transform Infrared Spectroscopy. New
York: Wiley.
Madsen, L. L., Helms, K., Zaba, B. N., Underbill, A.
E., Van der Sluijs, M. I (1991), Synth. Met. 43,
2931.
Magonov, S. N. (1993), Appl. Spectrosc. Rev. 28, 1.
Magnonov, S. N., Bar, G., Gorerberg, A. Y. (1993 a),
Advanced Materials 5, 453.
Magnonov, S. N., Seiko, S. S., Deblieck, R., Moller,
M. (1993 b), Macromolecules 26, 1380.
Mallik, R. R., Pritchard, R. G., Horley, C. C ,
Comyn, J. (1985), Polymer 26, 551.
Mansfield, T., Stein, R. S., Composto, R. X,
Rafailovich, M. H., Sokolov, X, Schwarz, S. A.
(1991), Physica B 173, 207.
Matienzo, L. X, Emmi, R, Egitto, F. D., van Hart, D.
C. (1988), /. Vac. ScL Technol. A6, 950.
Mayo, C. S., Hallock, R. B. (1989), Rev. ScL Instrum.
60, 739.
Marietta, G., Pignataro, S., Toth, A., Bertoti, I.,
Szekely, T., Keszler, B. (1991), Macromolecules 24,
99.
McGarvey, C. E., Holden, D. A., Tchir, M. F. (1991),
Langmuir 7, 2669.
McGlip, X F. (1990), J. Phys. Cond. Matter. 2, 7985.
Messerschmidt, R. G., Harthcock, M. A. (Eds.)
(1988), Infrared Microscopy, Practical Spectroscopy Series Vol. 6. New York: Marcel Dekker.
Meyer-Ilse, W. (1991), in: X-Ray Microscopy III:
Michette, A., Morrison, G., Buckley, C. (Eds).
Berlin: Springer-Verlag, pp. 284-289.
Michael, R., Stuik, D. (1987), Appl. Surf. ScL 28, 53.
Miller, D. R., Hahn, O. H., Bohn, P. W. (1987), Appl.
Spectrosc. 41, 245.
Mills, P. X, Green, P. R, Palmstrom, C. X, Mayer, X
W, Kramer, E. X (1984), Appl. Phys. Lett. 45, 957.
Mills, P. X, Palmstrom, C. X, Kramer, E. X (1986), /.
Mater. ScL 21, 1479.
Mirabella, F. M. (1988), Appl. Spectrosc. 42, 1258.
Mitchill, D. P., Sproule, G. I., Graham, M. X (1990),
Surf. Interface. Anal. 15, 487.
Monkman, A. P., Bloor, D., Stevens, G. C , Stevens,
X C. H., Wilson, P. (1989), Synth. Metals 29, E277.
Nakamoto, K. (1986), Infrared and Raman Spectra of
Inorganic and Coordination Compounds, 4th ed.
New York: Wiley.
Namavar, R, Budnick, X I. (1987), Nucl. Instrum.
Methods Phys. Res. B15, 285.
Nawaz, Z., Cataldi, T. R. I., Knall, X, Somekh, R.,
Pethica, X B. (1992), Surf ScL 265, 139.

751

Nefedov, V. D. (1988), X-ray Photoelectron Spectroscopy of Solid Surfaces. Utrecht, The Netherlands:
VSP.
Nicklow, R. M., Pomerantz, M., Segmuller, A.
(1981), Phys. Rev. B 23, 1081.
Niehuis, E., van Velzen, P. N. T., Lub, X, Heller, T.,
Benninghoven, A. (1989), Surf Interface Anal. 14,
135.
Niehus, H., Spitzl, R. (1991), Surf. Interface Anal. 17,
287.
Ninomiya, K., Hirai, Y, Momose, A., Aoki, S.,
Suzuki, K. (1991), J. Vac. ScL Technol. A9, 1244.
NIST (1989), X-ray Photoelectron Database, NIST
Standard Reference Database 20. Gaithersburg,
MD: NIST.
Nuzzo, R. G., Yang, Y H., Schwartz, G. P. (1987),
Langmuir 3, 1136-1140.
Nuzzo, R. G., Dubois, L. H., Allara, D. L. (1990). J.
Am. Chem. Soc. 112, 558.
Obrzut, X, Obrzut, M. X, Karasz, F. E. (1989), Synth.
Metals 29, E109.
Occhiello, E., Morra, M., Morini, G., Grarbassi, R,
Johnson, D. (1991), /. Appl. Polym. ScL 42, 2045.
O'Connor, D. X, MacDonald, R. X (1989), in: Ion
Beams for Materials Analysis: Bird, X R., Williams,
X S. (Eds.). New York: Academic Press.
Ohtani, H., Kao, C. T, Van Hove, M. A., Somorjai,
G. A. (1986), Prog. Surf ScL 23, 155.
Ohuchi, R S., Preilich, S. C. (1988), J. Vac. ScL Technol. A6, 1004.
Ong, T. H., Ward, R. N., Davies, P. B., Bain, C. D.
(1992), /. Am. Chem. Soc. 114, 6243.
Orti, E., Bredas, X L., Pireaux, X X, Ishihara, N.
(1990), J. Electron Spectrosc. Relat. Phenom. 52,
551.
Outka, D. A., Stohr, X (1988), J. Chem. Phys. 88,
3539.
Outka, D. A., Stohr, X, Rabe, X P., Swalen, X D.
(1988), /. Chem. Phys. 88, 4076.
Pachence, X M., Blasie, X K. (1987), Biophys. J. 52,
735.
Pan, R M., Lin, Y. L., Horng, S. R. (1991), Appl.
Surf ScL 47, 9.
Parikh, A. N., Allara, D. L. (1992), J. Chem. Phys. 96,
927.
Parikh, A. N., Sheen, C. W, Allara, D. L. (1993), in
press.
Paterson, P. X K. (1987), Mater. Forum 10, 144.
Pawson, D. X, Aneen, A. P., Short, R. D., Denison,
P., Jones, I. R. (1992), Surf. Interface Anal. 18, 13.
Payne, R. S., Clough, A. S., Murphy, P., Mills, P. X
(1989), Phys. Rev. B42, 130.
Pemberton, X E., Sobocinski, R. L., Bryant, M. A.,
Carter, D. A. (1990), Spectroscopy 5, 26.
Penfold, X, Thomas, R. K. (1990), J. Phys. Cond.
Matter. 2, 1369.
Perry, S. S., Campion, A. (1991), Surf ScL 259, 207.
Pertsin, A. L., Pashunin, Y. M. (1991), Appl. Surf. ScL
47, 115.

752

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

Pfeffer, R., Lux, R., Berkowitz, H., Lanford, W. A.,


Burman, C. (1982), / Appl. Phys. 53, 4226.
Pianetta, P., King, P. L., Borg, A., Kim, C , Lindau,
L., Knapp, C , Keenlyside, M., Browning, R.
(1990), J. Electron Spectrosc. Relat. Phenom. 52,
797.
Pireaux, J. X, Riga, X, Caudano, R., Verbist, X (1981),
in: Photon, Electron and Ion Probes of Polymer
Structures and Properties, ACS Symp. Series 162.
Washington DC: ACS.
Pireaux, X X, Thiry, P. A., Caudano, R., Pfluger, P.
(1986), J. Chem. Phys. 84, 6452.
Pireaux, X X, Gregoire, C , Vermeesch, M., Thiry, P.
A., Caudano, R. (1987), Surf. Sci. 189/190, 903.
Pireaux, X X, Vermeersch, M., Gregoire, G., Thiry, P.
A., Caudano, R., Clarke, T. C. (1988), J. Chem.
Phys. 88, 3353.
Pireaux, X X, Vermeersch, M., Degosserie, N., Gregoire, C , Novis, Y, Chtaib, M., Caudano, R.
(1989), Adhesion and Friction: Grunze, M.,
Kreuzer, H. X (Eds.). Berlin: Springer-Verlag.
Pireaux, X X, Riga, X, Boulanger, P., Snauwaert, P.,
Novis, Y, Chtaib, M., Gregoire, C, Fally, R,
Beelen, E., Caudano, R., Verbist, X (1990), /. Electron Spectrosc. Relat. Phenom. 52, 423.
Pireaux, X X, Thiry, P. A., Sporken, R., Caudano, R.
(1990a), Surf Interface Anal. 15, 189.
Pireaux, X X, Gregoire, C , Vermeersch, M., Thiry, P.
A., Vilar, M. R., Caudano, R. (1990b), in: Metallization of Polymers, ACS Symp. Series 440. Washington, DC: ACS.
Pireaux, X X, Gregoire, C , Caudano, R., Vilar, M.
R., Brinkhuis, R., Schouten, A. X (1991), Langmuir
7, 2433.
Pollard, X D., Sambles, X R. (1987), Optics Commun.
64, 529.
Porter, M. D. (1988), Anal. Chem. 60, 1143A.
Porter, M. D., Bright, T. B., Allara, D. L., Chidsey,
C. E. D. (1987), J. Am. Chem. Soc. 109, 3559.
Powell, C. X (1991), Surf. Interface Anal. 17, 308.
Powell, C. X, Seah, M. P. (1990), J. Vac. Sci. Technol.
A8, 735.
Powell, C. X, Seah, M. P. (1991), J. Vac. Sci. Technol.
A8, 735.
Purtell, R. X, Pomerantz, M. (1991), Langmuir 7,
2443.
Puydt, Y. D., Bertrand, P., Lutgen, P. (1988), Surf
Interface Anal. 12, 482.
Puydt, Y D., Leonard, D., Bertrand, P. (1990), in:
Metallization of Polymers, ACS Symp. Series 440.
Washington, DC: ACS.
Rabe, X P., Swalen, X D., Outka, D. A., Stohr, X
(1988), Thin Solid Films 159, 275.
Rabolt, X E, Schlotter, N. E., Swalen X D. (1981), /.
Phys. Chem. 85, 4141.
Rabolt, X F., Burns, F. C , Schlotter, N. E., Swalen,
X D. (1983), J. Chem. Phys. 78, 946.
Rafailovich, M. H., Sokolov, X, Jones, R. A. C ,
Krausch, G., Klein, X, Mills, R. (1988), Europhys.
Lett. 5, 657.

Rafailovich, M. H., Sokolov, X, Zhao, X., Jones, R.


A. L., Kramer, E. X (1990), Hyper fine Interactions
62, 45.
Ramaker, D. E. (1991), Crit. Rev. Solid State Mater.
Sci. 17,211.
Reimer, L. (1984), Transmission Electron Microscopy:
Physics of Image Formation of Micro analysis.
Berlin: Springer-Verlag.
Reiter, G., Bucbeck, C, Stamm, M. (1992), Langmuir
8, 1881.
Reiter, G., Huttenbach, S., Foster, M., Stamm, M.
(1991), Fresnius J. Anal. Chem. 341, 284.
Reilley, C. N., Everhart, D. S. (1982), in: Applied
Electron Spectroscopy for Chemical Analysis'. Windawi, H., Ho, F. F. L. (Eds.) New York: Wiley.
Reneker, D. H. (1990), in: New Characterization
Techniques for Thin Polymer Films: Tong, H. M.,
Nguyen, L. T. (Eds.). New York: Wiley.
Richmond, G. L., Robinson, X M., Shannon, V. L.
(1989), Prog. Surf Sci. 28, 1.
Ritsko, X X (1979), J. Chem. Phys. 70, 5343.
Ritsko, X X, Bigelow, R. W. (1978), /. Chem. Phys. 69,
4162.
Ritsko, X J, Brillson, L. X, Bigelow, R. W, Fabish, T.
X (1978), /. Chem. Phys. 69, 3931.
Roessler, A. (1987), Nauch. Appar. 2, 57; Mikrochim.
Ada 2, 79.
Rondelez, F , Aussere, D., Hervet, H. (1987), Ann.
Rev. Phys. Chem. 38, 317.
Rothberg, L., Higashi, G. S., Allara, D. L., Garoff, S.
(1987), Chem. Phys. Lett. 133, 67.
Rous, P. X, van Hove, M. A., Somojai, G. A. (1990),
Surf. Sci. 226, 15.
Russell, T. P. (1990), Mater. Sci. Rep. 5, 111.
Russell, T. P. (1991), Annu. Rev. Mater. Sci. 21, 249.
Russell, T. P., Karim, A., Mansour, A., Felcher, G. P.
(1988), Macromolecules 21, 1890.
Russell, T. P., Coulon, G., Deline, V R., Miller, D. C.
(1989), Macromolecules 22, 4600.
Russell, T. P., Menelle, A., Hamilton, W A., Smith,
G. S., Satija, S. K., Majkrzak, C. F (1991), Macromolecules 24, 5721.
Sauer, B. B., Walsh, D. X (1991), Macromolecules 24,
5948.
Salaneck, W R. (1981), in: Photon, Electron and Ion
Probes of Polymer Structures and Properties, ACS
Sympos. Series 162. Washington, DC: ACS.
Salaneck, W R. (1985), CRC Crit. Rev. Solid State
Mater. Sci. 12, 261.
Salaneck, W R. (1986), in: Handbook of Conducting
Polymers, Vol. 2: Skotheim, T. (Ed.). New York:
Marcel Dekker.
Salaneck, W R. (1991), Rep. Prog. Phys. 54, 1215.
Sarid, D. (1992), Scanning Force Microscopy, with
Applications to Electric, Magnetic, and Atomic
Forces. New York: Oxford University Press.
Sarid, D., Elings, V. (1991), /. Vac. Sci. Technol. B9,
431.
Schnatter, K. H., Doremus, R. H, Lanford, W A.
(1988), J. Non-Cryst. Solids 102, 11.

21.4 References

Schwamm, D., Kulig, J., Litt, M. H. (1991), Chem.


Mater. J, 616.
Scoles, G. (Ed.) (1992), Atomic and Molecular Beam
Methods, Vol. 2. New York: Oxford University
Press.
Seah, M. P. (1980), Surf. Interface Anal. 2, 222.
Seah, M. P. (1984), Vacuum 34, 463.
Seah, M. P. (1986), Vacuum 36, 399.
Seah, M. P. (1991), J. Vac. Sci. Technol A9, 1227.
Seah, M. P., Briggs, D. (1990), in: Practical Surface
Analysis by Auger Electron and X-ray Photoelectron Spectroscopy: Briggs, D., Seah, M. P. (Eds.).
New York: Wiley.
Seah, M. P., Dench, W. A. (1979), Surf. Interface
Anal. 1, 2.
Seah, M. P., Smith, G. C. (1988), Surf Interface Anal.
11, 69.
Selvam, P., Viswanathan, B., Srinvasan, V. (1990),
Electron Spectrosc. Relat. Phenom. 50, 277.
Senoh, T., Sanui, K., Ogata, N. (1990), Chem. Lett.,
1849.
Seul, M., Eisenberger, P., McConnell, H. M. (1983),
Proc. Natl. Acad. Sci. USA 80, 5795.
Shanker, K., MacDonald, J. R. (1987), /. Vac. Sci.
Technol. A5, 2984.
Shannon, C , Campion, A. (1988), J. Phys. Chem. 92,
1385.
Sheer, W, Allara, D. L. (1993), unpublished results.
Shen, Y. R. (1984), The Principles of Nonlinear Optics.
New York: Wiley.
Shen, Y. R. (1989), Annu. Rev. Phys. Chem. 40, 327.
Sherwood, P. M. A. (1991), /. Vac. Sci. Technol. A9,
1493.
Shull, K. R., Kramer, E. J., Hadziiounou, G., Antoietti, M., Sillesiu, H. (1988), Macromolecules 21,
2587.
Shull, K. R., Kramer, E. J., Hadziioannou, G., Tang,
W (1990), Macromolecules 23, 4780.
Shull, K. R., Dai, K. H., Kramer, E. X, Fetters, L. X,
Anloietti. M., Sillecu, H. (1991 a), Macromolecules
24, 505.
Shull, K. R., Kramer, E. X, Bates, F. S., Rosedale, X
H. (1991b), Macromolecules 24, 1383.
Skull, K. R., Winey, K. L, Thomas, E. L., Kramer, E.
X (1991c), Macromolecules 24, 2748.
Siegbahn, K. (1986), Phil. Trans. Royal Soc. Lond.
A318, 3.
Siegbahn, K. (1990), J. Electron Spectrosc. Relat.
Phenom. 51, 11.
Siegbahn, K., Nordling, C , Johansson, G., Hedman,
X, Heden P. E, Hamrin, K., Gelius, U., Bergmark,
T, Werme, L. O., Manne, R. (1969), ESCA Applied
to Free Molecules. Amsterdam: North-Holland.
Skotheim, T. A., Yang, X. Q., Chen, X, Inagaki, T,
Boer, M. D., Tripathy, S., Samuelsen, L., Rubner,
M. E, Hong, K., Watanabe, I., Okamoto, Y
(1989), Thin Solid Films 178, 233.
Smith, R., Harrison, D. E., Garrison, B. X (1990),
Nucl. Instrum. Methods Phys. Res. 46, 1.

753

Snyder, S. R., White, H. S. (1992), Anal. Chem. 64.


Sokolov, X, Rafailovich, M. H., Jones, R. A. L.,
Kramer, E. X (1989), Appl Phys. Lett. 62, 280.
Sokolov, X, Rafailovich, M. H., Jones, R. A. L.,
Kramer, E. X (1989 a), Appl. Phys. Lett. 54, 590.
Solomon, X L., Madix, R. X, Stohr, X (1990), J. Chem.
Phys. 93, 8379.
Sondag, A. H. M., Raas, M. C , Touwslager, F. X
(1991), Appl. Surface Sci. 47, 205.
Stamm, M. (1992), in: Advances in Polymer Science,
Vol. 100: Macromolecules: Synthesis, Order and
Advanced Properties, to be published.
Stamm, M., Majkrzak, C. F. (1987), Polymer
Preprints 28, 18.
Stamm, M., Reiter, G., Kunz, K. (1991), Physica B
173, 35.
Stamm, M., Huttenbach, S., Reiter, G., Springer, T.
(1991a), Europhys. Lett. 14, 451.
Stamm, M., Reiter, G., Kunz, K. (1991 b), Physica B
173, 35.
Steiner, U., Krausch, G., Schatz, G., Klein, X (1990),
Phys. Rev. Lett. 64, 1119.
Stickle, W. E, Moulder, X E (1991), / Vac. Sci. Technol. A9, 1441.
Stocker, W, Bickmann, B., Lotz, B., Moller, M.,
Magnonov, S. N., Cantow, H. X (1992), Ultramicroscopy 42144, 1141.
Stohr, X, Outka, D. (1987), /. Vac. Sci. Technol. A5,
919.
Stole, S. M., Popenoe, D. D., Porter, M. D. (1991),
Electrochemical Interfaces: Modern Techniques for
in-situ Interface Characterization. VCH, 339-410.
Storp, S., Holm, R. (1979), J. Electron Spectrosc. Relat. Phenom. 16, 183.
Stranick, S. X, Weiss, P. S., Parikh, A. N., Allara, D.
L. (1993), J. Vac. Sci. Technol, in press.
Stranick, S. X, Weiss, P. S. (1993), Rev. Sci. Instrum.,
in press.
Strohmeier, B. R. (1991), Appl. Surg. Sci. 47, 225.
Strohmeier, B. R., Leger, D. E., Field, R. S., Hercules, D. M. (1985), J. Catalysis 94, 514.
Strong, L., Whitesides, G. M. (1988), Langmuir 4,
540.
Strydom, I. L., Hofmann, S. (1991), J. Electron Spectrosc. Relat. Phenom. 56, 85.
Takutaka, H., Nishimori, K., Matomura, X, Pinet, V
(1987), Surf. Sci. 186, 339.
Thomas, S., Sherwood, P. M. A., Lee, K. M., Shea,
M. X O. (1990), Chem. Mater. 2, 7.
Tidswell, I. M., Rabedeau, T. A., Pershan, P. S.,
Folkers, X P., Baker, M. V, Whitesides, G. M.
(1991), Phys. Rev. B. 44, 869.
Tillman, N., Ulman, A., Penner, T. (1989), Langmuir
5, 101.
Tippman-Krayer, P. T., Mohwald, H. (1991), Langmuir 7, 2303.
Tsai, W. H., Young, X T, Boerio, E X (1991a), /.
Adhesion. 33, 211.
Tsai, W H., Boerio, E X, Clarson, S. X, Parsonage, E.
E., Tirrell, M. (1991b), Macromolecules 24, 2538.

754

21 Characterization of Surfaces, Interfaces, and Thin Films of Organic Materials

Tsai, W. H., Boerio, F. I, Jackson, K. M. (1992),


Langmuir 8, 1443.
Turner, N. H. (1988), Anal. Chem. 69, 377R.
Turner, N. H. (1990), Anal. Chem. 62, 113R.
Turner, N. H., Schreifels, J. A. (1992), Anal. Chem.
64, 302R.
Ulman, A. (1991), An Introduction to Ultrathin Organic Films from LangmuirBlodgett to Self-Assembly. New York: Academic Press.
Umemura, X, Kamata, T., Kawai, T., Takenaka, T.
(1990), /. Phys. Chem. 94, 62.
Vargo, T. G., Gardella Jr., J. A., J. Polym. Sci. Part A
(1989), Polym. Chem. 27, 1267.
Valenty, S. I, Chera, J. X, Smith, G. A., Katz, W,
Argain, R., Bakhru, H. (1984), /. Polym. Sci.
Polym. Chem. 22, 3367.
van der Wei, H., Lub, X, van Velzen, P. N. T.,
Beninghoven, A. (1990), Mikrochim. Acta II, 3.
van Ooij, W. X, Michael, R. S. (1990), in: Metallization of Polymers, ACS Sympos. Series 440. Washington, DC: ACS.
van Ooij, W. X, Sabata, A., Appethans, A. D. (1991),
Surf. Interface Anal. 17, 403.
Vasile, M. X, Bachman, B. X (1989), /. Vac. Sci. Technol. A7, 2992.
Venkatesan, T., Wolf, T., Allara, D., Wilkens, B. X,
Taylor, G. N. (1983), Appl. Phys. Lett. 43, 934.
Venkatachalam, R. S., Boerio, F. X, Roth, P. G., Tsai,
W H. (1988), /. Polym. Sci. Polym. Phys. Ed. 26,
2447.
Vilar, M. R., Schott, M., Pireaux, X X, Gregoire, C ,
Caudano, R., Lapp, A., Silva, X L., Rego, A. M.
(1989), Surf Sci. 2111212, 782.
Vilar, M. R., Scholt, M., Pireaux, X X, Gregoire, C ,
Thiry, P. A., Caudano, R., Lapp, A., Rego, A. M.,
Silva, X L. (1987), Surf. Sci. 189/190, 927.
Vogel, V., Shen, Y. R. (1991), Annu. Rev. Mater. Sci.
21, 515.
Vogel, V, Woll, C. (1988), Thin Solid Films 159, 429.
Wagner, C. D. (1972), Anal. Chem. 44, 1050.
Wagner, C. D. (1977), Anal. Chem. 49, 1282.
Wagner, C. D., Zatko, D. A., Ramond, R. H. (1980),
Anal. Chem. 52, 1445.
Walczak, M. M., Chung, C , Stole, S. M., Widrig, C ,
Porter, M. D. (1991), /. Am. Chem. Soc. 113, 2370.
Walls, D. X (1991), Appl. Spectrosc. 45, 1193.
Wandass, X H., Gardella Jr., X A. (1987), Langmuir 3,
183.
Wandass, X H., Gardella Jr., X A. (1986), Langmuir 2,
543.
Ward, R. X, Wood, B. X (1992), Surf Interface Anal.
18, 679.
Wasserman, S. R., Whitesides, G. M., Tidswell, I. M.,
Ocko, B. M., Pershan, P. S., Axe, X D. (1988), J.
Am. Chem. Soc. Ill, 5852.
Wasserman, S., Whitesides, G. M., Tidswell, I. M.,
Ocko, M., Pershan, P. S., Axe, X D. (1989), /. Am.
Chem. Soc. Ill, 5882.
Watts, X F. (1988), Surf Interface Anal. 12, 497.

Whitlow, S. X, Wool, R. P. (1989), Macromolecules


22, 2648.
Williams, D. B. (1984), Practical Analytical Electron
Microscopy in Materials Science. Mahwah, NJ:
Philips Electronic Instruments.
Williams, D. E., Davis, L. E. (1977), in: Characterization of Metal and Polymer Surfaces, Vol. 2; Polymer Surfaces'. Lee, L. H. (Ed.). New York: Academic Press.
Willis, R. F. (1990), J. Electron Spectrosc. Relat. Phenom. 54/55, 445.
Winograd, N. (1993), Anal. Chem. 65, 622A.
Wolf, S. G., Lindau, E. M., Lahav, M., Leiserowitz,
L., Deutsch, M., Kjaer, K., Als-Nielsen, X (1988),
Thin Solid Films 159, 29.
Woodruff, D. P., Delchar, T. A. (1986), Modern Techniques of Surface Analysis. New York: Cambridge
University Press.
Woollam, X A., Snyder, P. G. (1990), Mater. Sci. Eng.
B 5, 279.
Xie, X Z., Murarka, S. P., Guo, X. S., Lanford, W A.
(1988), J. Vac. Sci. Technol. B6, 1756.
Xue, G., Zhang, X F. (1991), Appl. Spectrosc. 45, 760.
Xue, G., Dong, X, Zhang, X F. (1991), Macromolecules 24, 4195.
Yates, X T., Madey, T. E. (Eds.) (1987), Vibrational,
Spectroscopy of Molecules on Surface. Vol. 1:
Methods of Surface Characterization. New York:
Plenum.
Yeh, P. (1988), Optical Waves in Layered Media. New
York: Wiley.
Zhang, Y, Levy, Y, Lougergue, X C. (1987), Surf. Sci.
184, 214.
Zhang, X Y, Shen, Y. R., Soane, D. S. (1992), J. Appl.
Phys. 71, 2655.
Zhao, W, Zhao, X., Rafailovich, M. H., Sokolov, X,
Mansfield, T., Stein, R. S., Composto, R. X,
Kramer, E. X, Jones, R. A. L., Sansone, M., Nelson, M. (1991), Physica B 173, 43.
Zhao, X., Zhao, W, Rafailovich, M. H., Sokolov, X,
Russell, T. P., Kumar, S. K., Schwarz, S. A.,
Wilkens, B. X (1991a), Europhys. Lett. 15, 725.
Zimba, C. G., Hallmark, V. M., Turrell, S., Swalen, X
D., Rabolt, X F. (1990), /. Phys. Chem. 94, 939.

General Reading
Andrade, X (Ed.) (1985), Surface and Interface Aspects of Biomedical Polymers. New York: Plenum
press.
Bird, X R., Williams, X S. (Eds.) (1989), Ion Beams for
Materials Analysis. New York: Academic Press.
Brewis, D. M. (1982), Surface Analysis and Pretreatment of Plastics and Metals. New York: Macmillan.
Brewis, D. M., Briggs, D. (1985), Industrial Adhesion
Problems. New York: John Wiley.

21.4 References

Briggs, D., Seah, M. P. (Eds.) (1991), Practicel Surface Analysis by Auger and X-ray Photoelectron
Spectroscopy, 2nd Ed. New York: John Wiley.
Chu, W. K., Mayer, J. W, Nicolet, M. A. (1978),
Backscattering Spectrometry. New York: Academic Press.
Feldman, L., Mayer, J. W. (1986), Fundamentals of
Surface and Thin Film Analysis. Amsterdam: North
Holland.
Howie, A., Vuldre, U. (Eds.) (1988), Surface and Interface Characterization by Electron Optical Methods. New York: Plenum Press.
Ibach, H., Mills, D. L. (1982), Electron Energy Loss
Spectroscopy and Surface Vibrations. New York:
Academic Press.
Neagle, W., Randell, D. R. (Eds.) (1990), Surface
Analysis Techniques and Applications. London:
Royal Society of Chemistry.

755

Nefedov, V. D. (1988), X-ray Photoelectron Spectroscopy of Solid Surfaces. Utrecht, The Netherlands:
VSP.
Riviere, J. C. (1990), Surface Analysis Techniques.
New York: Oxford University Press.
Tong, H. M., Nguyen, U. T. (Eds.) (1990), New Characterization Techniques for Thin Polymer Films,
New York: John Wiley.
Windami, H., Ho, F. F. L. (Eds.) (1982), Applied Electron Spectroscopy for Chemical Analysis. New
York: John Wiley.
Woodruff, D. P., Delchar, T. A. (1989), Modern Techniques of Surface Science. Cambridge: University
Press.
Wu, S. H. (1982), Polymer Interfaces and Adhesion.
New York: Marcel Dekker.

Potrebbero piacerti anche