You are on page 1of 21

Vibrational spectroscopy of interfaces by infrared–visible sum frequency generation

M. Buck and M. Himmelhaus

Citation: Journal of Vacuum Science & Technology A 19, 2717 (2001); doi: 10.1116/1.1414120
View online: http://dx.doi.org/10.1116/1.1414120
View Table of Contents: http://scitation.aip.org/content/avs/journal/jvsta/19/6?ver=pdfcov
Published by the AVS: Science & Technology of Materials, Interfaces, and Processing

Articles you may be interested in


Understanding rubber friction in the presence of water using sum-frequency generation spectroscopy
J. Chem. Phys. 130, 024702 (2009); 10.1063/1.3049582

Conformational changes at polymer gel interfaces upon saturation with various liquids studied by infrared-visible
sum frequency generation vibrational spectroscopy
Appl. Phys. Lett. 88, 134105 (2006); 10.1063/1.2188388

Structure of the glycerol liquid/vapor interface studied by sum-frequency vibrational spectroscopy


Appl. Phys. Lett. 84, 4965 (2004); 10.1063/1.1762699

Effect of a static electric field on the vibrational and electronic properties of a compressed CO adlayer on Pt(110)
in nonaqueous electrolyte as probed by infrared reflection–absorption spectroscopy and infrared-visible sum-
frequency generation spectroscopy
J. Chem. Phys. 119, 12492 (2003); 10.1063/1.1626640

Infrared and infrared-visible sum frequency generation spectroscopic response of harmonic monolayer vibrons:
Homogeneous profile
J. Chem. Phys. 110, 6963 (1999); 10.1063/1.478602

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
REVIEW ARTICLE

Vibrational spectroscopy of interfaces by infrared–visible sum


frequency generation
M. Bucka)
School of Chemistry, University of St. Andrews, North Haugh, St. Andrews KY16 9ST, United Kingdom
M. Himmelhausb)
Lehrstuhl für Angewandte Physikalische Chemie, Ruprecht Karls Universität Heidelberg, INF 253,
69120 Heidelberg, Germany
共Received 31 August 2001; accepted 4 September 2001兲
During the past decade vibrational sum frequency generation as a method to study interfaces has
matured and can now be applied more routinely to systems of increasing complexity. The article
provides a brief overview of technical aspects of infrared–visible sum frequency generation,
compares this nonlinear technique with its linear analogs, and highlights the latest applications.
© 2001 American Vacuum Society. 关DOI: 10.1116/1.1414120兴

I. INTRODUCTION ments. A level of reliability and quality of SFG spectro-


meters has now been reached which allows a more wide-
Despite the diversity of issues related to buried interfaces,
spread application to more complex systems in various fields
one faces the general problem that investigation of them re-
of technological importance such as catalysis, chemical va-
quires analytical techniques which have to be very sensitive,
por deposition, electrochemistry, and even topics related to
i.e., detecting molecules at a level below 1015 per cm2 , to be
very specific to yield the required information, and, at the environmental problems or life science. This development is
same time, to be applicable under nonvacuum conditions. reflected by a significant increase in the number of groups
The latter constraint imposes the most severe restriction starting SFG and by the commercialization of complete SFG
since techniques based on strongly interacting, massive par- spectrometers.
ticles like electrons or ions cannot be applied and, conse- A number of reviews, which cover the fields SFG has
quently, the zoo of available techniques is greatly reduced. substantially contributed to throughout the past decade, have
Unfortunately, the remaining techniques more or less lack at been published and the impressive range of topics addressed
least one of the other requirements, and, thus, limit the level highlights the versatility of this technique.2–14 The range ex-
of information about interfaces or the range of systems they tends from interfaces of liquids,7,8 the properties of
can be applied to. surfactants,3,4 catalytic reactions,10,15 and electro-
Fortunately, a technique which combines all three require- chemistry,6,16 to the dynamics at interfaces.9
ments to a substantial degree emerged in 1987. Providing Considering the number of reviews available, what could
access to interfaces by being photon based, having an intrin- another one add to that not discussed before? From numer-
sically high interface sensitivity due to the nonlinear interac- ous discussions with colleagues, who are unfamiliar with but
tion of light with matter, and being specific by probing mo- are interested in this technique, we got the impression that, in
lecular vibrations, sum frequency generation 共SFG兲 from a most cases, there are no clear thoughts of SFG beyond the
monolayer of coumarin 504 on fused quartz was attained by well-known interface sensitivity. Since linear optical vibra-
the group of Shen1 by combining a visible with an infrared tional spectroscopies are familiar techniques, we believe a
beam 共IR–vis SFG兲. However, widespread application of comparison of them with SFG might be beneficial to appre-
SFG in the subsequent years was hampered by significant ciate the potential of SFG but, equally important, also to
disadvantages. A rather limited tunability, the time spent on become aware of limitations and pitfalls. For those who are
maintenance and optimization of the laser system, together
more seriously considering SFG as a technique to start with,
with the high costs restricted SFG as an analytical tool to a
a discussion of different options of SFG spectrometers might
few specialized groups rather than to nonexpert users. Even
also be helpful. Trivially, a review should provide an update
though it still is not a cheap technique, the situation has
of latest applications and comment on directions. Accord-
fundamentally changed throughout the past decade due to
tremendous progress in laser technology which has removed ingly, the article is organized such that after a brief summary
major technical obstacles to performing IR–vis SFG experi- of basic theoretical aspects in Sec. II, some design principles
and common layouts of SFG spectrometers are discussed
a兲
Electronic mail: mb45@st-and.ac.uk 共Sec. III兲, and a more detailed comparison of SFG with other
b兲
Electronic mail: il7@ix.urz.uni-heidelberg.de optical vibrational techniques is provided, in particular IR

2717 J. Vac. Sci. Technol. A 19„6…, NovÕDec 2001 0734-2101Õ2001Õ19„6…Õ2717Õ20Õ$18.00 ©2001 American Vacuum Society 2717

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2718 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2718

共DFG兲. In the case of degeneracy, i.e., ␻ 1 ⫽ ␻ 2 , SHG and


static polarization, called optical rectification, occur. Since
what is to be presented equally holds for DFG and SFG, we
only refer to SFG further on. The superscript denoting the
order of the susceptibility is also omitted.
To illustrate sum frequency generation at interfaces we
take a typical geometry, sketched in Fig. 1. Two incoming
waves with wave vectors k 1i and k 2i approach the interface
between a linear 共1兲 and a nonlinear medium 共2兲 whose op-
tical properties are represented by ⑀ 1 and ⑀ 2 , respectively.
The optical properties of the interfacial region may differ
from ⑀ 1 and ⑀ 2 and, therefore, an interfacial dielectric con-
stant ⑀ I is introduced. Governed by the linear optical prop-
erties of the layered system the waves are partially reflected
(k 1/2r ) and transmitted (k 1/2t ). Correspondingly, sum fre-
quency waves generated in medium 2 propagate along the
directions of reflection (k SF,r ) and transmission (k SF,t ). The
FIG. 1. Schematic of SFG at interfaces. The incoming beams with wave
vectors k 1i and k 2i impinge on the surface and generate surface polarization
directions of the SF signals are determined by momentum
P I which gives rise to a SFG signal with wave vectors k SF,r/t . In the case of conservation parallel to the interface:
higher order contributions or a noncentrosymmetric medium an additional
background due to a bulk polarization P B adds coherently to the pure sur- k SFx ⫽k 1x ⫹k 2x . 共4兲
face signal. Instead of measuring the reflected signal the SFG signal gener-
ated in transmission can also be used. As pointed out in detail by Shen,21 the source term of the
sum frequency signal P( ␻ SF) can be separated into an inter-
face contribution PI and a bulk term PB generated inside
spectroscopy 共Sec. IV兲. Some selected recent examples are medium 2:
presented in Sec. V followed by some concluding, critical
remarks. P共 ␻ 兲 ⫽PI 共 ␻ 兲 ⫹PB 共 ␻ 兲 . 共5兲

In general, P is dominated by the electric dipole contribution


II. THEORY of the bulk term, since the magnitude of the nonlinear cou-
Within the framework of this article the following presen- pling drops with increasing multipole order and bulk contri-
tation of the theoretical background has, necessarily, to be butions dominate over surface contributions of the same
superficial. Nevertheless it covers the essential aspects. Thor- order.18 However, if medium 2 has a center of inversion a
ough descriptions of the fundamentals of nonlinear optics, in dipole contribution is forbidden by symmetry. Therefore,
general,17–20 and surface nonlinear optics, in particular, are only the dipole contribution from the interface, where the
available.2– 4,21,22 Articles covering the latter include second inversion symmetry is broken, and higher order terms from
harmonic generation 共SHG兲 which has the same physical the bulk are left. This means that the interface signal can be
background and preceded SFG. of the same order as or dominate over the bulk contributions.
If an external electric field 共E兲 becomes comparable to the It is important to note that the origin of PI typically is on the
intramolecular field felt by an electron, the linear polariza- order of a few angströms across the interface. This is in
tion of matter is no longer a sufficient approximation and contrast to most linear optical methods used for interface
higher order terms have to be taken into account. The total studies such as infrared reflection absorption spectroscopy
polarization 共P兲, using the electric dipole approximation, is 共IRRAS兲, ellipsometry, or methods which rely on evanescent
then described by a series expansion fields. At best their surface sensitivity addresses a regime on
the order of the wavelength of the light, i.e., several hun-
P⫽ ␹ 共 1 兲 E⫹ ␹ 共 2 兲 :EE⫹ ␹ 共 3 兲 ]EEE....., 共1兲 dreds of nanometers or even micrometers. Therefore, second
where the linear and nonlinear properties of the material are order optical techniques provide a unique tool with which to
represented by the susceptibilities ␹ (n) . For two light waves probe the immediate interfacial region. Even though methods
of different frequencies ␻ 1 and ␻ 2 , and amplitudes E 0,1 and that utilize surface enhancement of the electromagnetic fields
E 0,2 the first nonlinear term gives rise to mixing, are also quite surface sensitive due to a rapid decay of the
field, they have severe restrictions. Surface enhanced Raman
P共 2 兲 ⫽ ␹ 共 2 兲 :E0,1 cos共 ␻ 1 t 兲 E0,2 cos共 ␻ 2 t 兲 , 共2兲 spectroscopy 共SERS兲, for example, requires rough surfaces
and yields, by applying trigonometric identities, and is limited to a small number of substrate materials.
Before analyzing the interface signal in more detail, we
P共 2 兲 ⫽ 21 ␹ 共 2 兲 :E0,1E0,2关 cos共 ␻ 1 ⫹ ␻ 2 兲 t⫹cos共 ␻ 1 ⫺ ␻ 2 兲 t 兴 . 共3兲
rewrite the second order polarization as
These polarization components which oscillate at the sum
and the difference frequencies of the two incident waves are
the origins of SFG and difference frequency generation
P⫽ 兺
i jk
ê i ␹ i jk E j E k 共6兲

J. Vac. Sci. Technol. A, Vol. 19, No. 6, NovÕDec 2001

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2719 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2719

the adsorbate itself which, here, are meant to originate from


all internal and external molecular vibrations ( ␹ v ) have to be
taken into account. In referring to the literature for
details21,29 we note that ␹ sub and ␹ int can also contain contri-
butions of higher multipoles some of which cannot be sepa-
rated from the dipole term. Furthermore, ␹ sub incorporates
any bulk contributions from the substrate. Usually, these
complications are resolved by subsumming ␹ sub and ␹ int un-
der a single susceptibility ␹ eff which depends on the adsor-
bate coverage ␪. We rewrite Eq. 共7兲 as

I⬀ 冏兺 冉
i jk
K eff,i jk ␹ eff,i jk 共 ␪ 兲 ⫹ 兺␯ K ␯ ,i jk ␹ ␯ ,i jk 冊冏 2
,

with ␹ eff,i jk 共 ␪ 兲 ⫽ ␹ sub,i jk ⫹ ␹ int,i jk 共 ␪ 兲 , 共8兲

where the K’s are a contracted notation of the corresponding


quantities in Eq. 共7兲. For a typical SFG experiment which
FIG. 2. Illustration of the sources of a sum frequency signal from an adsor- combines a beam at fixed frequency in the visible or near IR
bate. Besides the resonant susceptibilities from the molecular vibrations with a beam tunable in the range of the molecular vibrations,
共both internal and external兲 which are represented by the susceptibilities ␹ ␯ , the substrate signal usually can be considered nonresonant,
the substrate can give rise to a nonresonant signal from the surface elec- i.e., it does not change appreciably within the spectral range
tronic polarizability ( ␹ sub) which in addition can be altered by the formation
of the adsorbate substrate bond ( ␹ int).
investigated. Even though the case of an adsorbate will be
mostly discussed throughout the present article Eq. 共8兲 can
be straigthforwardly applied to describe SFG from surfaces
of bulk samples such as polymers or liquids. In these cases
to point out that the susceptibilities are third rank tensors ␹ eff and ␹ v represent the properties of the bulk and the mo-
with, in general, 27 nonvanishing elements. Here, ê i denotes lecular moieties at the surface, respectively.
the unit polarization vector of component i. The number of Besides providing the vibrational fingerprint of an inter-
nonzero elements of ␹ is usually greatly reduced and de- face, SFG allows one to determine the orientation of mol-
pends on the symmetry of the system.18 For example, a sur- ecules. This requires relating the nonlinear properties of the
face with azimuthal isotropy (C ⬁ v ) is described by only four molecule, described by the hyperpolarizability ␤, with the
independent elements ␹ zzz , ␹ zxx ⫽ ␹ zy y , ␹ xzx ⫽ ␹ yzy , and experimentally accessible, macroscopic susceptibility ␹ and,
␹ xxz ⫽ ␹ y yz . Furthermore, by choosing proper polarization thus, involves the transformation of ␤ from the molecular
combinations of the incoming beams and the analyzed SFG coordinate system (abc) to the surface coordinate system
signal, e.g., ss p, 23 even single tensor elements can be probed i jk, i.e., ␤ abc ⇒ ␤ i jk . This scheme is based on a number of
and this way the orientation of molecules can be restrictions and assumptions such as the averaging over
analyzed4,24 –26 as detailed later. For other polarizations such many molecules, i.e., an orientational distribution, no inter-
as ppp the situation is more complicated since different ten- action between the molecular fragments, i.e., ␹ ⫽N 具 ␤ 典 with
sor elements contribute and, therefore, complicate the inter- N as the number density of a SFG active moiety, and pertur-
pretation of the SFG signal. Accounting for the tensor char- bation theory being appropriate to derive ␤. In this case Eq.
acter of ␹ the intensity of the sum frequency signal is given 共8兲 can be rewritten as
by

I⫽C ␻ SF
2
冏兺 兺
i jk m
ê i K m,i ␹ m,i jk K m, j K m,k E 0j 共 ␻ 1 兲 E 0k 共 ␻ 2 兲 冏 2
; I 共 ␻ SF兲 ⬀ 冏兺i jk
K eff,i jk ␹ eff,i jk 共 ␪ 兲 ⫹N 兺␯ K ␯ ,i jk 具 ␤ ␯ ,i jk 典 冏 2
,

共7兲 共9兲

C is a constant which depends on details of the setup7,18 and with


i jk refers to a Cartesian coordinate system. Note, that in Eq.
共7兲 the superscript 0 denote the fields at the beam source, in ␣ ␯ ,i j ␮ ␯ ,k
contrast to in Eq. 共6兲, where the field components are the ␤ ␯ ,i jk ⫽ , 共10兲
␻ IR⫺ ␻ ␯ ⫹i⌫ ␯
local ones. The correction for the local field is then taken into
account by the K’s which comprise the Fresnel factors and where ␻ v and ⌫ v are the resonance frequency of mode ␯ and
local field effects.18,27,28 The summation over m indicates that its bandwidth. Equation 共10兲 shows that ␤ is proportional to
the SFG signal can be a coherent superposition from differ- both the Raman and IR transition dipole moments4,6,30 whose
ent sources as illustrated in Fig. 2 for an adsorbate/substrate components are represented by ␣ v ,i j and ␮ v ,k , respectively.
system. Contributions from the bare substrate ( ␹ sub), its al- The form of Eq. 共10兲 describes a singly resonant transition
teration due to the bonding of an adsorbate ( ␹ int), and from in the IR by a Lorentzian profile, thus neglecting inhomoge-

JVST A - Vacuum, Surfaces, and Films

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2720 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2720

hence, the molecular orientation. This is a problem which


has been treated in the literature rather diversely. In most
cases, the linear optical properties of the adsorbate layer are
simply neglected completely. For small adsorbate molecules
such as CO on Pt this is a satisfactory solution which, how-
ever, might be problematic in the case of more complex sys-
tems such as ultrathin polymer films or self-assembling
monolayers.
The analysis of the orientation is also affected by the
components of the Raman tensor ␣, a second rank tensor.
Even though in special cases, such as in vibration of B 1
FIG. 3. Schematic of the SFG process which involves IR and Raman tran- symmetry in a molecule of C 2 v symmetry, one can derive the
sitions. The dashed line indicates a virtual level in the Raman transition. molecular orientation by probing different tensor elements
However, this could also be a real one and it would give further resonance without knowing ␤,4,38 other cases require the tensor ele-
enhancement of the SFG signal.
ments of ␣ to be known.4,24,33 In principle, they can be cal-
culated by ab initio techniques but reliable accurate calcula-
tion of them is a problem which has not been solved in
neous broadening. Even though such a line profile seems too general. Also, only relatively few values based on experi-
simplistic it has been a satisfying approximation for most mental data exist and even those can vary substantially.39
purposes, possibly due to the relatively large linewidth of Consequently, given the number of assumptions and uncer-
most laser systems used for SFG. There are only a few ex-
tainties in the input data required to extract orientational in-
amples given in the literature that also account for inhomo-
formation from SFG spectra, an accurate determination of
geneous effects on the line shape.31–33 The product form of
molecular orientation remains a challenging task as has been
the transition dipole moment and Raman tensor in Eq. 共10兲
detailed by Bell et al..33
shows that in SFG both IR and Raman transitions are in-
As mentioned above, ␹ ␯ is of dipolar origin and thus van-
volved, as schematically depicted in Fig. 3. This also reflects
ishes in centrosymmetric media. Therefore, the SFG re-
that ␹ ␯ vanishes in centrosymmetric media because from
sponse of a certain vibrational mode ␯ contains information
group theory it is well known that for a molecule exhibiting
about the distribution of the corresponding molecular subunit
a center of inversion a respective vibrational mode is either
within the adsorbate layer. The resonance can disappear in a
IR or Raman active.
SFG spectrum if the molecular entity is either isotropically
A few remarks are appropriate here. In general, both the
susceptibilities and the K’s are complex quantities and, there- distributed or exhibits centrosymmetric crystalline packing.
fore, due to the coherent nature of frequency mixing a SFG Besides its high interface specificity, this sensitivity to struc-
spectrum is crucially dependent on their mutual phase rela- tural and conformational order is one of the major reasons
tion. Depending on their relative phase, the interference be- for the success of IR–vis SFG in surface analysis, since it
tween resonant contributions as defined by Eq. 共9兲 and the provides information about adsorbate layers that is hard to
background signal ( ␹ eff) can range from constructive to de- gain by any other method. Numerous examples of this prop-
structive. Furthermore, if the K’s are complex and if more erty in different systems such as ultrathin organic
than one tensor element of a particular vibrational mode con- films,30,31,40,41 polymers,42– 44 and ice/water interfaces45,46
tributes, the phase of the resulting contribution from this have been reported and an illustrative example will be pre-
mode depends on the orientation of the respective molecular sented in Sec. IV.
moiety. This further complicates the evaluation and an accu- For systems with a non-negligible signal from the sub-
rate description requires a separate phase factor for each strate surface, an additional source of information unique to
mode.34 the nonlinear techniques can be used to trace an adsorbate
If the phase relations are known the average orientation of coverage. As indicated in Eq. 共8兲 by the interaction term
the molecular entities can, in principle, be derived from mea- ␹ int( ␪ ), the nonresonant substrate signal yields information
surements in different polarization combinations.4,24 –26 Hi- about the coverage and, thus, can be used to monitor a cov-
rose et al. have provided a detailed analysis of SFG spectra erage dependent evolution of molecular orientation and con-
of methylene and methyl groups which, in principle, can be formation. For chemisorbed adsorbates ␪ dependent changes
transferred to other molecular systems.25,26 After deduction of ␹ eff are easily detected.40,47,48 However, a quantitative
of the hyperpolarizability ␤ of a respective molecular unit, evaluation requires caution. First, it cannot be assumed a
the transformation from the molecular to the laboratory co- priori that the relationship between the coverage and the
ordinate system is performed by assuming a certain distribu- change of the magnitude of ␹ int( ␪ ) is linear. Second, the
tion function for the molecular entities.35 This leads to a phase can significantly change with ␪ and, thus, can alter the
direct correlation between the macroscopic strength of the shape of the spectrum. Considering that molecular reorienta-
resonance and the molecular orientation. That way, structural tion can be coverage dependent and, as mentioned above,
information is obtained.36,37 However, from Eq. 共9兲 it is seen can also produce a phase shift, the separation of both effects
that proper K’s are crucial for the determination of 具␤典 and, can become difficult. Third, the intensity of the resonances

J. Vac. Sci. Technol. A, Vol. 19, No. 6, NovÕDec 2001

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2721 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2721

relative to the nonresonant background changes and this has a tuning range from 4500 to 1000 cm⫺1 at sufficiently high
to be taken into account in the evaluation. These points output power 共typically 10–100 ␮J/pulse in the ps regime兲. A
which have been studied in detail by SHG49 but are directly very critical point of a SFG spectrometer is the pump source
applicable to SFG make it mandatory for each system to whose performance is crucial for the noise, long term stabil-
calibrate the coverage dependence of the nonresonant SFG ity, and reliability of a SFG experiment. There are flash lamp
signal with another independent surface sensitive technique. pump lasers based on saturable dyes, but they cause rather
Similarly, caution is also required in electrochemical experi- large energy fluctuations and have relatively poor beam pro-
ments when interpreting potential dependent data quantita- files. The introduction of new modelocking techniques,68 –73
tively since a potential change can also affect the substrate regenerative amplifiers, and diode pumping has substantially
signal.6 Finally, for constant coverage, the substrate contri- improved the situation.
bution to the total signal is usually assumed to be nonreso- Depending on the techniques the temporal sequence of
nant within the tuning range of the IR radiation. This is not pulses is different. Most benchtop systems generate pulses
always clear in advance.29 For example, for silver surfaces it with equal temporal spacing where the repetition rate ranges
has been found experimentally as well as theoretically that from the tens of Hz region 共see e.g., Refs. 40, 44, and 74兲 to
they exhibit a pronounced resonance when probed in the the hundreds of Hz/kHz range.52,59,75 In another scheme
range of the CH stretching vibrations.50 bunches of pulses are produced. Within a single bunch the
repetition rate is around 100 MHz whereas the bunch fre-
III. EXPERIMENTAL ASPECTS quency is in the ten Hz range. Such a scheme is used with
FELs and also in a benchtop system. The latter has the ad-
A. Laser systems
vantage that by using a parametric oscillator rather than para-
The success and increasing application of IR–vis SFG are metric generators and amplifiers, tunable IR down to 10 ␮m
intimately linked to advancements in laser technology and, is achieved in a single step.68 Since the different timings of
despite the diversity of possible technical realizations of a the laser pulses can impose rather different transient thermal
SFG spectrometer, the generation of tunable IR radiation loads on the systems investigated, the temporal structure of
with intensities sufficiently high for surface SFG can be the pulse sequence is something to keep in mind. At this
identified as their common bottleneck. To achieve such high point it might be interesting to take a brief look at the factors
intensities pulsed sources are required and until rather re- which determine the performance of a SFG spectrometer. As
cently this was restricted to pulses in the nano- to picosecond a general measure of performance, let us define W SF as the
regime where IR pulses with a spectral width of a few wave average power of the sum frequency signal. With increasing
numbers are generated and the SFG spectrum is acquired by W SF the signal to noise 共S/N兲 ratio is improved and thus the
scanning over the spectral range of interest.6 Another, more acquisition time can be reduced, yielding better performance
recent, scheme allows one to simultaneously record a spec- of the system. Then, the energy per single pulse, w SF , can be
trum. Introduced by van der Ham et al.51 and by Richter written as
et al.52 using a free electron laser 共FEL兲 and a benchtop fem-
tosecond 共fs兲 system, respectively, it relies on a broadband W SF
w SF⫽ , 共11兲
IR source. For example, a Fourier limited broad band IR n ␯ rep
pulse in the 100 fs range which corresponds to a range of
where n denotes the number of pulses in a single bunch and
about 300 cm⫺1 is combined with a visible pulse. The latter
␯ rep the repetition rate of the bunches. For simplicity we
has a spectral width of a few wave numbers and, thus, pro-
consider temporal and spatial beam profiles of rectangular
vides the spectral resolution.52–54
shape that match perfectly, i.e., ␶ SF⫽␶vis⫽␶IR and A SF⫽A vis
Among the schemes to generate tunable IR, nano- and
⫽A IR , with ␶ being the pulse duration and A the projected
picosecond pulses, Raman shifters,31,55–57 difference fre-
beam diameter. Then, from Eq. 共11兲 it becomes clear that
quency generation in nonlinear optical crystals,30,52,58 – 61
parametric conversion,62 and FELs63,64 have mostly been w SF SF w visw IR
used. There is a clear dominance of frequency mixing in ⫽I ⬀ 兩 P SF兩 2 ⬀I visI IR⫽ 2 2 ,
␶A ␶ A
nonlinear crystals, even though Raman shifters, which can be
共12兲
run efficiently with nanosecond pulses, have been success- w visw IR
fully applied.3,44,65 FELs provide high output powers over w SF⬀ ,
␶A
the widest tuning range of all the techniques. They have
shifted the low frequency limits of SFG down to less than where w vis and w IR denote the pulse energies of the incoming
500 cm⫺1,66 although the successful generation of IR down beams. This expression demonstrates an intrinsic property of
to 500 cm⫺1 by frequency mixing has also been reported.67 nonlinear optical 共NLO兲 processes, i.e., both shortening of
Unfortunately, FELs, being large scale facilities, impose re- the pulse duration and focusing increase the number of SFG
strictions on SFG experiments such as continuous availabil- photons per pulse and, thus, improve the S/N ratio of the SF
ity and, thus, benchtop laser systems are clearly to be pre- signal. This is the reason for using pulsed lasers in NLO
ferred whenever possible. Today’s tabletop systems based on spectroscopy and for the tendency to reduce the pulse dura-
IR generation by parametric conversion and DFG are the tion as much as possible. However, from a practical point of
best choice for the majority of applications. They now cover view there are some limitations to this strategy. First of all, a

JVST A - Vacuum, Surfaces, and Films

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2722 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2722

reduction in the pulse duration increases the natural line-


width of the pulses. For transform limited pulses the line-
width of pulses of a few ps duration is on the same order as
the bandwidth of the vibrational modes which, for adsorbates
is typically in the range of several wave numbers. As long as
one is only interested in spectroscopic characterization and
not in any dynamics, further reduction of the pulse duration
thus does not result in further improvement of the resonant
SF signal, because the linewidth of the pulse is spread be-
yond the resonant regime of an individual resonance. There-
fore, a pulse duration ␶ of a few ps seems to give the best
trade-off between high conversion efficiency and the best FIG. 4. Schematic of experimental geometries used in SFG. 共a兲 Normal
reflection and transmission, 共b兲 total internal reflection, 共c兲 counterpropagat-
excitation of resonant features.76 Of the laser systems run- ing geometry, and 共d兲 surface plasmon enhanced SFG.
ning today there is large variation in the performance de-
pending on the technology.6 The example of a SFG spectrum
of a thiol self-assembled monolayer on gold might illustrate will usually be determined by the pumping source, e.g.,
this. At comparable S/N ratio and spectral range, a spectrum around 800 nm for Ti–sapphire systems and 532 nm/1064
takes about 1 min with a fs system operating at 1 kHz,52 nm for Nd–YAG, unless parametric conversion of the fixed
whereas a 20 Hz picosecond system requires about 1 h.40 frequency allows tuning.
Another implication is related to the question of to what
extent can the projected beam diameter A be reduced. Here, B. Geometry
the damage threshold of the surface comes into play. Damage
Similar to linear vibrational spectroscopies a number of
can occur due to either thermal effects or ablation. While
geometries which are depicted in Fig. 4 exist for SFG at
thermal effects like surface melting and boiling are the main
interfaces. The optimum geometry with respect to a maxi-
sources of damage for pulses longer than 100 ps, ultrashort
mum SF signal is determined by factors which are intrinsi-
pulses less 20 ps in duration mainly cause ablative effects,
cally related to both the linear optical properties of the inter-
such as photoionization and plasma formation.77 Thus, to
face and the nonlinearity of the SFG process. Mostly,
avoid surface damage the power density of an individual
reflection geometry 关Fig. 4共a兲兴 is used which, trivially, has to
pulse must not exceed the threshold for ablation, while the
be applied if one of the media is opaque. Furthermore, in the
average power deposited in the surface should be kept below
case of a contribution from a nonlinear bulk polarization P B
any significant thermal effects. This can be a serious issue 共see Fig. 1兲 this configuration is, in general, superior to trans-
and a check is mandatory for each system. In studies of mission geometry. The coherence length for SFG in reflec-
surface phase transitions the repetition rate had to be reduced tion is much shorter than in transmission which limits the
to eliminate artifacts78 and in SFG from Langmuir–Blodgett bulk SF generation to a volume very close to the interface
films the IR beam was found to induce disorder.37 The de- region, whereas in transmission geometry the signal can be
pendence of thermal effects on the pulse rate demonstrates generated along a distance which is of the order of funda-
that accumulative effects of subsequent pulses occur. This mental wavelengths.29 Nevertheless, in cases where the indi-
can become a more serious issue for experiments using pulse ces of refraction of the adjacent media are close to each other
bunches although it strongly depends on the thermal proper- the reflected signal is small and transmission geometry might
ties of the system under study. As a guideline one should be preferred. If both the reflected and transmitted SF signals
keep in mind that the thermal load deposited in a surface by can be measured the ratio of surface and bulk contributions
a single pulse dissipates on a time scale of ns for metals and can be determined.29,74
doped semiconductors, but can be in the range of micro- to Among the reflection geometries one can differentiate be-
milliseconds for insulators. tween normal 关Fig. 4共a兲兴 and total internal reflection 共TIR兲
Another point is the choice of the visible wavelength. 关Fig. 4共b兲兴. The latter requires that the indices of refraction
Since the intensity of the SFG signal scales with ␻ 2 关see Eq. on the beam side are larger than those on the opposite side.
共7兲兴,7,79 the number of photons as re-radiated by the nonlin- Furthermore, a prism geometry has to be chosen in order to
ear polarization P will vary like ␻. Therefore, ␭ SFG and thus allow the angle of incidence ␣ i of beam i to satisfy the con-
␭ vis should be minimized to optimize the S/N ratio. However, dition of total reflection. The Fresnel factors can be maxi-
other factors might oppose this and force a compromise. Ad- mized if the ␣ i ’s are close to the critical angles ␣ i,T . 38,81– 85
justment of ␭ vis to the system under investigation might be According to Eq. 共7兲 this increases the SFG signal compared
the most desirable way to minimize the nonresonant back- to that in normal reflection geometry and an enhancement of
ground and/or to increase the damage thresholds as, for ex- more than an order of magnitude can be easily achieved.
ample, has been shown using ␭ vis⫽532 and 1064 nm.80 Dif- Two points concerning TIR have to be noted. First, there
ficulties in the spatial separation of the vis and SFG beams is a strong variation of the Fresnel factors around the angles
can also be factors for preferring longer wavelengths. of incidence. Consequently, small changes of the index of
However, in practice the wavelength of the visible light refraction alter ␣ i,T and, therefore, the Fresnel factors. A

J. Vac. Sci. Technol. A, Vol. 19, No. 6, NovÕDec 2001

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2723 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2723

good example was reported recently for a polymer film on a beam direction is required for this particular setup. As an
solid substrate.44 Since the indices of refraction of the sub- example, a scan within the range of the CH-stretching vibra-
strate and the polymer are significantly different, the ␣ i,T ’s of tions 共3050–2750 cm⫺1兲 with the visible and IR beams im-
the air/polymer and polymer/substrate interface differ. Ad- pinging on an air/solid interface at 50° and 70°, respectively,
justing the angles of incidence close to ␣ i,T of the air/ yields a change in angle of 0.19° for copropagating and 19.0°
polymer and the polymer/substrate interfaces, respectively, for counterpropagating geometries, respectively. Another im-
the molecular orientation at the two interfaces could be stud- portant point to note is that the spectra can depend on the
ied separately. However, the sensitivity of the Fresnel factors geometry. From a theoretical point of view this difference
to the index of refraction can also give rise to distortions of just corresponds to inversion of the in-plane component of
the SFG spectra since the index of refraction changes appre- the IR beam. For certain tensor elements this causes inver-
ciably within a resonance.7,85 Second, one has to keep in sion of the sign, i.e., a phase shift by ␲, and therefore, the
mind that the argument of field enhancement is only valid if SFG signal which consists of a coherent superposition of
the intensities at the interface are not limited by the optical contributions from different tensor elements changes its
damage threshold. If the damage threshold is limiting it does spectral shape.90
not matter by which geometry the limit for the interfacial As one last point in Sec. III we address the issue of dif-
electric fields is reached.7 However, the laser fluence can ference frequency generation. Besides the lower power emis-
always be adjusted below the threshold by irradiating a sion of DFG generated from the same pump frequencies, due
larger area. This enables more molecules to be sampled and, to ␻ DFG⬍␻SFG mentioned above 关Eq. 共7兲兴, the basic theory
thus, increases the signal. of nonlinear optics91 does not predict any difference in ap-
Another geometry, but one restricted to metal surfaces, is plicability between sum ( ␻ 3 ⫽ ␻ 2 ⫹ ␻ 1 ) and difference fre-
related to the TIR in the sense that it seeks an increase of the quency generation ( ␻ 3 ⫽ ␻ 2 ⫺ ␻ 1 ). However, up to now,
electric field at the interface. By exciting surface plasmon DFG has hardly been applied. One major reason might be
polaritons 共SPPs兲 the surface electric field can be strongly that upconversion of the visible light to the sum frequency
enhanced to yield a SFG signal that is amplified by orders of produces a SFG signal which is located in a spectral range
magnitude compared to in linear reflection geometry. To ex- that is free of fluorescence and, hence, in a regime in which
cite a SPP, the component of the k vector of the incoming noise is very low. In contrast, DFG is located in a range
where fluorescence due to the visible pump beam can inter-
electromagnetic wave parallel to the surface has to match the
fere and spoil the nonlinear signal. In recent
k vector of the SPP. In referring to the literature for details it
experimental92–95 as well as theoretical work96,97 it was
was reported86 that this requires special conditions such as
shown that it might be worth compensating for this disad-
thin semitransparent metal films or rough surfaces, e.g., grat-
vantage by improving the detection system. For cases in
ings 关Fig. 4共c兲兴. As discussed by Alieva et al.87 for excitation
which the SFG photon energy is located above the excitation
of SPPs by IR radiation, the site of SPP generation and that
level of the substrate and thus exhibits increased noise and
of NLO mixing can be separated due to the small longitudi-
nonresonant background, in particular, DFG can be an el-
nal damping constant of SPPs in the IR. Unfortunately, this
egant alternative. This was theoretically analyzed by Men-
scheme is only of limited practical value since the resonance
doza et al. for adsorbates on gold and compared to data of
bandwidth for a given angle of incidence is too narrow to be
the CN⫺/Au system.96 In this particular case the nonresonant
kept for the entire tuning range of the IR. Recently, Eliel and signal was significantly reduced for DFG compared to for
co-workers88,89 discussed several geometries for excitation of SFG, resulting in increased sensitivity to adsorbate reso-
SPPs and could demonstrate that the SFG response from the nance.
surface can be improved by a factor of 102 – 104 . Besides the
geometry depicted in Fig. 4共a兲 where the IR and vis beams
copropagate, a counterpropagating arrangement is also ap- IV. COMPARISON WITH OTHER VIBRATIONAL
plied frequently 关Fig. 4共d兲兴. According to the conservation of SPECTROSCOPIES
momentum parallel to the surface 关Eq. 共4兲兴 this increases the As already discussed in Sec. I, considering the significant
angle between the sum frequency and the visible beams and, experimental effort, limitations in tunability and costs of
thus, facilitates their spatial separation. This becomes in- SFG, alternative photon-based vibrational spectroscopies
creasingly important the longer the IR wavelength or the such as IR and Raman should be carefully considered before
shorter the visible wavelength is. For example, to improve deciding to apply SFG. Following from the discussion in
the ratio of the SFG signal to the background from the high Sec. II three key points are identified in which the physical
intensity of the visible beam, counterpropagating geometry is principles of the techniques manifest. First, for surface stud-
favorable for measurements in the long wavelength range ies of materials whose bulk exhibits a vibrational structure in
共⬎10 ␮m兲. On the other hand, counterpropagation geometry the same spectral range as the surface species, SFG has an
and the bigger influence of the IR beam on the direction of obvious advantage. Typical examples of this type are sur-
the SFG signal can also be a disadvantage, since scanning faces of polymers or liquids. Second, the diversity of spectral
the frequency results in angular dispersion of the SFG signal features is expected to be reduced for SFG compared to its
which is larger than in copropagating geometry. Therefore, a linear optical counterparts, since ␮ and ␣ determine the num-
detection system with a large aperture or correction of the ber of observable vibrations 关Eq. 共9兲兴 which often are mutu-

JVST A - Vacuum, Surfaces, and Films

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2724 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2724

ally excluding. However, this argument can be also turned tion of the CO coverage evolved differently. The divergent
around since a small cross section for either ␮ or ␣ can be results were explained by coverage dependent changes of the
compensated for by a large one of the other and, therefore, Raman tensor. For adsorbates on metal substrates, several
vibrational features suffering from very low intensity in one additional points can make quantitative interpretation of the
of the linear techniques might be raised to a significant level SFG more difficult compared to IRRAS. The first one is the
in SFG. No conclusive statement on this point can be pro- lack of complete screening of the parallel components of the
vided yet. The systems investigated so far, which were ex- visible light, i.e., the components of ␣ parallel to the sub-
tensively investigated in the frequency range above 1600 strate, known in Raman spectroscopy as the breakdown of
cm⫺1 共see Table I兲, have both significant IR and Raman ac- the selection rule.108 However, in practice the extent of
tivity. For the lower frequency range where the interesting screening is strongly dependent on the metal and on the vis-
functional group frequencies are located the question re- ible wavelength used. The second one is that the phase rela-
mains unresolved but experiments on selected examples tion is substrate dependent and, furthermore, depends on the
point towards the first case.98 The third point which defi- orientation of a molecular entity. A third factor is that for
nitely still is a disadvantage of SFG is concerned with the adsorbate–substrate interactions strong enough to signifi-
tunability. IR and Raman spectroscopies allow simultaneous cantly perturb the polarizability of the substrate, the substrate
detection of the whole spectral range whereas SFG is a scan- contribution changes as a function of coverage. The latter
ning technique which, in addition, becomes increasingly dif- two points were already addressed in Sec. II.
ficult towards longer wavelength, i.e., ⬎1000 cm⫺1. This There are a number of differences between SFG and IR-
still holds for the broadband detection scheme with fs pulses RAS which provide valuable complementary information.
共Sec. III A兲 since the wavelength range covered simulta- The most striking one is the sensitivity of SFG to molecular
neously is in the range of 200–300 cm⫺1. A typical acquisi- symmetry as exemplified in Fig. 5 which shows a compari-
tion time of an IRRAS spectrum is a few minutes for a range son of the IRRAS and SFG spectra of a complete self-
of typically 600– 4000 cm⫺1 whereas for SFG a 250 cm⫺1 assembled monolayer of docosane thiol
range needs on the order of 1–2 min.52 关 H3C共CH2兲21SH,MC22兴 on a polycrystalline, 共111兲-textured
Raman spectroscopy has a decided advantage over SFG gold film.40 The upper and the lower parts of Fig. 5 display
共and IR spectroscopy兲 in that it operates in a spectral range in the SFG and IRRAS spectra, respectively. The middle spec-
which many materials are transparent. Furthermore, it con- trum is a deconvolution of the SFG spectrum based on Eq.
veniently allows microscopy. Unfortunately, the low sensitiv- 共9兲 assuming Lorentzian profiles. In the IRRAS spectrum the
ity is a severe limitation and it has prevented Raman spec- band intensities largely reflect the relative quantities of CH3
troscopy from becoming a general analytical technique for and CH2 and, thus, the spectrum is dominated by the meth-
adsorbate studies. This, of course, does not mean that, apart ylene vibrations. In contrast, the SFG spectrum strongly em-
from surface enhanced Raman scattering,99 monolayers can- phasizes the methyl vibrations whereas the methylene bands
not be investigated. Under favorable conditions, that is, if the are essentially absent. It is known that alkane thiols form
scattering cross section is high enough like that for hydrogen densely packed monolayers with the hydrocarbon chains in
on Si共111兲,100,101 if resonant Raman scattering102 can be an all-trans conformation109 and obviously SFG stresses the
used, or if the electric field at the interface can be enhanced CH3 groups which terminate the self-assembled monolayer
by plasmon excitation103 or total reflection geometry,82,104 共SAM兲 and which are arranged in a strongly noncentrosym-
the required sensitivity can be achieved. However, all these metric fashion. Conversely, the absence of the CH2 signal
cases impose restrictions on the choice of system and to date indicates that the methylene moieties are present in a cen-
Raman spectroscopy in its unenhanced version is a technique trosymmetric structure, in full agreement with the all-trans
that is applied only seldom to study adsorbates. chain of the model. One detail which further illustrates the
This is not the case for IR spectroscopy which is a mature, sensitivity of SFG to centrosymmetry is the small feature
powerful technique105,106 and among all surface science tech- labeled d ⫹
t which is assigned to the methylene group adja-
niques one of the methods used most. For this reason a more cent to the methyl moiety and, thus has an asymmetric
detailed comparison of IR spectroscopy with SFG is given to environment.34,110 The situation is reversed for the same sys-
illustrate their complementarity and their specific advantages tem in a disordered state which, now, yields a strong SFG
or drawbacks beyond the points already addressed in Sec. II. signal from the methylene modes. This is shown in Fig. 6 for
Looking again at Eqs. 共9兲 and 共10兲 it is obvious that the a submonolayer coverage of MC22 which was prepared by
interpretation of SFG data is more complex compared to immersion of the substrate into a ␮molar thiol solution for
IRRAS. One reason is that the product of ␮ and ␣ enters into only a few minutes.40 The broad feature between 2900 and
the SFG signal. In addition to problems related to ␮ which 2920 cm⫺1 arises from the asymmetric and Fermi resonances
IRRAS and SFG share and which are caused by changes of of the symmetric CH2 stretching mode and indicates that the
the transition dipole moments due to intermolecular cou- chains are in a conformationally disordered state. The ab-
pling, the polarizability ␣ adds to this and further compli- sence of the CH3 signals, which is expected for an isotropic
cates the situation. An illustrative example which highlights orientation of the methyl groups, confirms the model of a
this problem is a comparative study of CO on Ni共111兲,107 highly disordered layer. Note that for this particular system
where the dependence of IRRAS and SFG signals as a func- there is another difference between IRRAS and SFG. There

J. Vac. Sci. Technol. A, Vol. 19, No. 6, NovÕDec 2001

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2725 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2725

TABLE I. List of systems investigated by SFG. Experiments of dynamics and those at the electrochemical interface are denoted by 共d兲 and 共ec兲, respectively,
and sol⫽solid, liq⫽liquid, vac⫽vacuum, and vap⫽vapor.

Type of
Species/spectral range 共cm⫺1兲 System/substrate interface Reference

OH stretch, 3000–3900

H2O H2O liq/vap 32, 160, 161


Hexane liq/liq 162, 163
Pt共111兲 sol/vac 46, 164
Quartz sol/liq 162
Surfactant, CCl4 liq/liq, vap 8, 161, 163, 165–169
Acids and salts liq/vap 170–179
Alcohols liq/vap 115, 180–183
Alkylnitrile liq/vap 181, 184, 185
Polymer film sol/liq 42

N–H stretch, 3000–3450


NH3 H2O liq/vap 186

CH stretch, 2700–3100
CH3OH CH3OH liq/vap 134
Polymer
Polymers sol/air 42– 44, 158, 187–192, 286
Water sol/liq
Sapphire
Self-assembled monolayers
Thiols Metal sol/air, liq 40, 41, 55, 80, 193–195,
90共ec兲, 3, 31, 85, 196 –199
Siloxanes Silica, CaF2 , glass sol/air, liq, sol 141, 200–203共d兲, 204共d兲

Amphiphiles
Langmuir films Water, glass, quartz, sapphire, sol/air, sol 39, 54, 205共d兲, 37, 82, 141, 142, 145, 147
metal liq/air 206, 207共d兲, 208共d兲
Surfactants H2O liq/air 36, 110, 209–213
30, 33, 78, 206, 214 –216
CCl4 liq/liq 84, 135, 83
Metal, quartz, glass sol/liq 117, 119, 217–220
Alcohols Alcohols liq/vap 182
Methane sulfonic acid H2O liq/vap 176, 221
CNCH3 Pt共111兲 sol/liq 共ec兲 118
ZrO2 sol/liq, gas 65
H, Cx Hy Diamond C(hkl) sol/vac, gas 130–133, 222, 223共d兲

Heterogeneous catalysis
Ethylidene Rh共111兲 sol/vac 200
Propionic acid Ni共110兲 sol/vac 224, 225
Formic acid NiO共111兲 sol/vac 151共d兲, 226共d兲
Ni共110兲 sol/vac 227, 228
Pt共110兲 sol/vac 229, 230
MgO共001兲 sol/vac 231, 232
Isobutene Pt共111兲 sol/gas 233
Propylene Pt共111兲 sol/gas 234
Cyclohexadiene Pt共111兲 sol/gas 235, 236
Cyclohexene Pt共111兲 sol/gas 237
Ethylene Pt共111兲 sol/gas 126, 238
Acetylene Pt共111兲 sol/gas 239

O–D stretch, 2100–2900


D2O Mica sol/liq 45
Formic acid NiO共111兲 sol/vac 150共d兲

C–D stretch, 2000–2300


Formic acid Ni共111兲, NiO共111兲 sol/vac 152
CD3共CH2兲19CN Langmuir film of air/water 24
Octadecane thiol Au sol/liq 3
Hydrogen Diamond C共111兲 sol/vac 132

CO stretch, 1700–2300
CO CO/polycrystalline Pt sol/gas 128, 240, 241
CO/Pt(hkl) sol/liq 共ec兲 242–247共d兲

JVST A - Vacuum, Surfaces, and Films

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2726 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2726

TABLE I. Continued

Type of
Species/spectral range 共cm⫺1兲 System/substrate interface Reference

CO/Pt(hkl) sol/gas, 66, 127, 248, 249


vac
CO/Pt(hkl) sol/liq 共ec兲 63, 122, 250, 251
CO/Cu共100兲 sol/vac 252共d兲
CO/Ru共100兲 sol/vac 153共d兲
CO/W共110兲 sol/vac 253
CO/Ni共100兲 sol/vac 254
CO/Ni共111兲,NiO共111兲 sol/vac 255共d兲, 107, 256, 257
CO/Pd共111兲 sol/vac 258, 259
CO/Pd nanoparticles sol/gas 75
CO/Pt nanoparticles sol/gas 129

NO stretch, 1500–2100
NO NO/Ni共111兲,NiO共111兲 257

Si–H stretch, 2070–2160 NO/Pt共111兲 260


Hydrogen Si共111兲 sol/vac 共149, 261–265兲共d兲, 266
Si共100兲/H sol/vac 267

H–metal, stretch, 1600–2300


Hydrogen Pt(hkl), polycrystalline sol/liq 共ec兲 268 –270

CN stretch, 1700–2300
CN Pt(hkl) sol/liq 共ec兲 251, 271–275共d兲
Ag sol/liq 共ec兲 246
Au(hkl) sol/liq 共ec兲 276, 277, also DFG
Pt, polycrystalline sol/liq 共ec兲 278
CH3CN ZrO2 sol/liq, gas 65
CD3共CH2兲19CN Langmuir film on H2O liq/vap 24
4-cyanopyridine Au共111兲 sol/liq 共ec兲 279
5-cyano-terphenyl H2O liq/vap 39
SCN
SCN/Pt sol/liq 共ec兲 278
SCN/Au sol/air, liq 85
OCN
OCN/Ag sol/liq 共ec兲 280

CvO, 1500–1850
Formic acid Ni共111兲 sol/vac 150
Polyimide Polyimide sol/air 191

C–C, 1300–1550
C60 Ag共111兲 62, 87, 281–283
2⫻1 surface Diamond C共111兲 sol/vac 132

C–H bending, 1300–1500


H Diamond C共111兲 sol/vac 132

Ar–NO2 , 1300–1350
Thiol Au sol/liq, air 284

Ar–C–C, 1000–1100
Thiophenol Ag sol/air 195

Aromatic ring deformations,


850–1300
Aromatic thiols Ag, Au sol/air 195, 285
4-cyanopyridine Au共111兲 sol/liq 共ec兲 279

C–S, 700/420
Thiophenol Ag sol/air 195

is a broad, intense peak at 2815 cm⫺1 in the SFG spectrum fore, are significantly redshifted. The difference in intensity
that is hardly discernible in IRRAS. It reflects methylene of this band between IRRAS and SFG can be explained by a
vibrations which interact with the metal surface and, there- large Raman cross section that can compensate for a weak IR

J. Vac. Sci. Technol. A, Vol. 19, No. 6, NovÕDec 2001

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2727 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2727

FIG. 5. Vibrational spectra of a complete self-assembled monolayer of FIG. 6. Vibrational spectra of an incomplete self-assembled monolayer of
docosane thiol on gold 共Ref. 40兲. 共a兲 The open circles in the SFG spectrum docosane thiol on gold 共Ref. 40兲. 共a兲 The open circles in the SFG spectrum
are the experimental data and the solid line is a fit based on Eq. 共9兲. 共b兲 are the experimental data and the solid line is a fit based on Eq. 共9兲. 共b兲
Vibrational susceptibilities normalized to the nonresonant substrate signal Vibrational susceptibilities normalized to the nonresonant substrate signal
␹ NR . Note the occurrence of upward and downward pointing bands which ␹ NR . The band assignments are the same as those in Fig. 5. Compared to
reflect the relative phases of the susceptibilities. 共c兲 IRRAS spectrum. Mode Fig. 5 note the additional feature labeled c S which is due to interaction of
assignments: d ⫹ symmetric CH2 , d ⫹ t symmetric mode of the CH2 group the hydrocarbon chain with the substrate.

adjacent to CH3 , r ⫹ symmetric CH3 , d ⫺ asymmetric CH2 , r FR Fermi reso-
nance of symmetric CH3 , r op asymmetric CH3 out of plane vibration, r ⫺

ip
asymmetric CH3 in-plane vibration.
␽ denotes the angle between the c axis and the surface nor-
mal. Assuming rotational symmetry around the c axis, ␣ cc is
the relevant element of the Raman tensor and r denotes the
transition dipole in SFG. Further, as seen from the analysis ratio of the Raman tensor elements ␣ aa / ␣ cc . The square
of the SFG either direction of the peaks can be measured if brackets denote the average value of the tilt angle. Referring
there is a nonresonant signal background 共here, from the sub- to the literature on how to determine molecular orientation
strate surface兲. from IRRAS111–113 and SFG25,30,35,36,114,115 we note that only
Upon changes of the polar orientation of molecules the in the case in which the distribution of tilt angles adopts a ␦
signal intensities for SFG and IRRAS are affected differ- function, i.e., the molecules have a well-defined orientation,
ently. As an illustrative example we consider the symmetric do IRRAS and SFG yield the same value. For other distribu-
methyl mode of a molecule adsorbed on a metal substrate. tions differing values are obtained and this provides informa-
Defining the transition dipole moment ␮ to be oriented along tion about the distribution.116 This, of course, holds for any
the c axis of an internal Cartesian coordinate system (a,b,c) comparison of SFG using a linear technique based on electric
the IR absorption for which only the component along the dipole transitions.
surface normal is relevant is given by Other cases of complementarity are those where both
techniques by themselves yield ambiguous results. Such
I z 共 IR兲 ⬀ 兩 ␮ c 具 cos ␽ 典 兩 2 , 共13兲
cases are changes from an oriented to an isotropic state and
whereas, e.g., for the zzz tensor element in SFG it is inversion of the orientation of molecules. In the first case, the
signal vanishes for SFG as discussed in Sec. II whereas IR
I zzz 共 SFG兲 ⬀ 兩 ␮ c ␣ cc 具 cos ␽ 典 关 1⫺ 共 1⫺r 兲 具 sin2 ␽ 典 兴 兩 2 ; 共14兲 still yields a signal. The other case of inversion leaves the IR

JVST A - Vacuum, Surfaces, and Films

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2728 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2728

unchanged when turning a molecule upside down whereas


the phase of the SFG signal changes by 180° and in the case
of interference with a nonresonant background the type of
interference changes and gives rise to a dramatic alteration of
the spectral shape. A good demonstration of this phase sen-
sitivity is identification of the polar orientation of a surfac-
tant described by Bain et al.117–119
For practical considerations a key advantage of SFG is
that it does not need sophisticated referencing since, even
though the efficiency is very low, it generates a new signal.
The only background which can affect the detection of reso-
nances in SFG comes from the nonresonant contribution of
the interface which varies from zero to about the same order
of magnitude as the resonance intensities 共if the media are
isotropic or centrosymmetric兲. IRRAS, in contrast, is a dif-
ference technique which either needs a clean substrate as a
reference or relies on a comparison of spectra acquired in s
and p polarization.120 Small changes of a large signal of the
order of 10⫺4 require normalization to a reference. This is
less of a problem for flat metal surfaces where the FIG. 7. Influence of the gas phase absorption on a nonresonant SFG signal
s-polarized signal serves as a reference due to screening of for CO. 共a兲 Gas phase absorption spectrum of CO together with the spectral
profile of the IR laser. The inset shows the CO spectrum over an extended
the in-plane field. However, for supported metal nanopar- range. 共b兲 Selected examples of the influence of the gas phase absorption on
ticles which, e.g., are of high interest for studies in catalysis, the SFG signal: 共i兲 undistorted signal; 共ii兲 spectrum at P CO⫽10⫺1 mbar, T
the selection rule can no longer be applied. Furthermore, for ⫽296 K; 共iii兲 P CO⫽10 mbar, T⫽296 K; 共iv兲 p CO⫽100 mbar, T⫽296 K;
transparent substrates, IRRAS becomes even more difficult 共v兲 P CO⫽100 mbar, T⫽600 K; 共vi兲 p CO⫽1000 mbar, T⫽296 K. In all cal-
culations the absorption length was set to 0.1 m. For clarity the spectra are
even though, by applying polarization modulation tech- displaced vertically. The peaks can easily reach 15% of the background
niques, spectra of rather good quality from monolayers on, signal. For further details see the text.
for example, liquid surfaces can be obtained.121 Another
clear advantage of SFG over IRRAS is the upconversion of
the IR into the vis/near-ultraviolet 共UV兲 range and the coher-
therefore, caution is required if the vibrational bands of the
ent nature of the process. Both allow the application of SFG adsorbate are in the same range as the gas phase species.
to high temperature processes where IRRAS can run into This is illustrated in Fig. 7 which shows a series of simulated
problems due to blackbody radiation and saturation of the SFG signals. Figure 7共a兲 depicts the gas phase absorption
detector. spectrum of CO together with the spectral profile of the IR
One of the often cited advantages of SFG is its possibility laser used in the simulations. A Lorentzian line shape was
to bridge the pressure gap in studies of surface reactions. assumed and the full width at half maximum 共FWHM兲 was
Even though this statement definitely holds, SFG is not un- set to 8 cm⫺1. A selection of calculated SFG signals is com-
affected by the bulk phase if IR absorption becomes notice- piled in Fig. 7共b兲. To demonstrate the influence of gas phase
able. This definitely imposes limitations to the investigation absorption a broad, featureless signal 共modeled by a Gauss-
of interfaces between condensed matter. As an example, the ian located at 2100 cm⫺1 with FWHM⫽500 cm⫺1兲 served as
interface between water and a solid can either be investigated the undistorted reference spectrum 关curve 共i兲兴. To study the
from the water side 共but this does not allow one to look at the effect of IR gas phase absorption on the SFG spectra the
interfacial water兲 or from the substrate side which, however, spectral profile of the IR laser was gas phase corrected. In the
severely limits the number of materials and/or the wave- first step, the gas phase spectra for the respective experimen-
length range. Solutions to minimize bulk absorption exist tal condition were calculated using Hitran database and
and electrochemical SFG experiments use the same cell de- software.123 Subsequently, the SFG signal given by the cen-
sign as IRRAS to reduce the layer thickness of the electro- ter wavelength of the laser line was calculated by summing
lyte solution122 to a few micrometers in order to limit bulk over the spectral profile of the laser line. As the resulting
absorption of the IR beam. spectra demonstrate, there is significant distortion of the SFG
Even though the particle density is lower by orders of spectra due to changes both in the pressure and in the tem-
magnitude, absorption in gaseous media can also appreciably perature which could easily be interpreted as surface species
affect a SFG spectrum and can be the source of artifacts. In generated by changes in the experimental conditions. This
fact, in this case things are more complicated than in con- means that SFG spectra have to be corrected for gas phase
densed matter. Since the absorption is caused by a large absorption and, thus, linear absorption spectra have to be
number of sharp rotational lines the light transmitted is recorded for all experimental conditions. Otherwise spectral
strongly structured. Furthermore, the band intensities of the changes can easily be misinterpreted as changes in the ad-
rotational lines are affected by temperature and pressure and, sorbate layer as was discussed recently for the CO/Pt共111兲

J. Vac. Sci. Technol. A, Vol. 19, No. 6, NovÕDec 2001

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2729 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2729

system.124 Two additional factors are important to consider.


First, for strong absorbers additional effects such as thermal
lensing125 can play a role and can alter the beam diameter on
the substrate. Since this causes additional changes in the
SFG signal such an influence of the gas phase is particularly
problematic because changes of this type cannot be corrected
by linear absorption measurements. Second, if reacting mix-
tures are investigated one has to be aware of the products and
concentration gradients, i.e., the gas phase may have to be
monitored along the path of the IR beam.

V. APPLICATIONS OF SFG: SELECTED EXAMPLES


Currently, IR–vis SFG spectroscopy is becoming a more
widespread technique and it can be expected that its range of
applications will be significantly extended in the future. So
far, as a look at Table I, which compiles the systems inves-
tigated to date, reveals there is a strong weighting both of the FIG. 8. SFG spectra of CO adsorbed on a Pd/Al2O3 model catalyst for two
wavelength range and of the systems studied. Investigations different Pd particle sizes 共upper spectrum: 6 nm; lower spectrum: 3 nm兲.
The pressure and the temperature were 10⫺7 mbar and 193 K, respectively.
in the shorter wavelength range 共⬎1600 cm⫺1兲 dominate
共Figure adapted from Ref. 75.兲
which, besides the fact that many interesting systems exhibit
vibrational features in this range, can be explained by its
easier access compared to the longer wavelength range. In-
terestingly this has not changed much over the years. In al- chemistry. This experiment marks an important, promising
most all experiments the range of either C–H stretching vi- step towards the investigation of models that closely ap-
brations of saturated and unsaturated hydrocarbons or small proximate commercial catalysts. Based on methods which
inorganic molecules such as CO, CN, NO, or H2O has been allow controlled preparation of nanoparticles on surfaces
investigated. With respect to the types of interface all pos- over a macroscopic area, the long sought after application of
sible combinations have been looked at. However, compared SFG to catalytic systems of technological relevance has be-
to other types of interfaces, reports on layers confined be- come feasible.75,129 Another point associated with studies of
tween two solids and liquid/liquid interfaces are very scarce. small particles is that SFG might have increased sensitivity.
For the latter the strong absorption of the IR for most mate- In an investigation of CO adsorbed on Pt nanoparticles en-
rials of interest is a serious obstacle to investigations of this hancement of the SFG signal from the CO by several orders
type of interface. of magnitude compared to that of flat surfaces was observed
In the vast majority of cases intramolecular vibrations and explained by plasmon enhancement.129 However, the ex-
were investigated. Very few SFG measurements have probed tension of SFG to study adsorbates on nanoparticles also
the adsorbate–substrate vibrations in the rather special cases means a step further in complexity and, consequently, inter-
of hydrogen on Si, metal, or diamond surfaces. Again, the pretation of the SFG signal becomes further complicated. A
frequency range that is accessible is responsible for this. The number of additional factors affect the signal such as the size
first successful attempt to probe adsorbate–substrate vibra- and the shape of the nanoparticles and, like in surface en-
tions of heavier particles that are well below 1000 cm⫺1 is hanced Raman scattering,99 one might face the problem of
yet to be made. An attempt to probe the Pt–CO vibration mixing of the contributions from chemical and field enhance-
共around 470 cm⫺1兲 was reported as unsuccessful.66 ment. The need for in situ detection under high pressure is
A field in which SFG has already been applied very suc- not only important for heterogeneous catalytic processes but
cessfully and in which it is expected to make further signifi- also for elucidation of the surface chemistry during thin film
cant contributions in the future is catalysis since it bridges growth by, e.g., chemical vapor deposition. One particularly
the pressure gap. Starting from UHV in the early days there demanding example is the low pressure synthesis of diamond
has been a move toward high pressures.15,124,126 –128 Because which takes place at high temperatures of the gas phase and
of its industrial importance, CO has been investigated exten- the substrate, and at pressures of several tens of millibars;
sively but other molecules of catalytic relevance such as conditions which have been prohibitive for other vibrational
small hydrocarbons, formic acid, or NO have also been stud- spectroscopies. Here, the coherent nature of SFG is a deci-
ied in detail. As a very recent example, which illustrates sive advantage since it allows spatial separation of the strong
current developments in this area, the CO oxidation on Pt but isotropic blackbody radiation from both the gas phase
clusters of two different sizes is shown in Fig. 8.75 The SFG and the substrate. Figure 9 shows SFG spectra of a C共111兲
spectra are characterized by two bands around 1980 and diamond surface which was exposed to an activated
2100 cm⫺1 which are assigned to bridge and on-top bonded hydrogen/methane gas mixture in a hot filament reactor. Un-
CO, respectively. Obviously, the size of the cluster pro- der deposition conditions which produce diamond, the SFG
foundly affects the bonding of the molecules and, thus, their spectrum is characterized by a pronounced peak at 2820

JVST A - Vacuum, Surfaces, and Films

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2730 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2730

FIG. 10. SFG spectra of a film of poly共ethylene glycol兲


关 H–共O–CH2 –CH2兲8 –OH) 共PEG400兲兴 on CaF2 . The film is in contact with
共a兲 ambient air and 共b兲 water. 共Figure adapted from Ref. 42.兲

such as CCl4 /H2O have been reported extensively in differ-


ent recent reviews.7,8,11,135 Thus, the present discussion is
limited to polymers whose surfaces are very interesting from
an application point of view since, for example, the control
of adhesion of polymers to each other and to other surfaces is
of utmost technological importance. The environment in-
duced change of the surface was demonstrated with a ure-
thane type molecule to which an polydimethylsiloxane chain
共PDMS兲 was grafted. Changing from an air to a water envi-
ronment the surface undergoes pronounced restructuring and
switches from hydrophobic to hydrophilic.43 The influence of
FIG. 9. Vibrational spectra of a diamond 共111兲 surface during hot filament the environment also holds for the polymer solid interface.
chemical vapor deposition. In the upper part a stream of a hydrogen/ As observed earlier by SHG which revealed a strong sub-
methane mixture passes a filament to produce atomic hydrogen and CH
species, leading to diamond growth. The total pressure is 30 mbar and the
strate dependence of the polymer film structure,136 SFG from
gas mixture is 0.5% CH4 in 99.5% H2 . The SFG spectra were acquired for a polystyrene 共PS兲 film on sapphire underlines the crucial
different substrate temperatures (T S ) and of the filament (T F ). The two sets influence of the adjacent medium. The aromatic moieties
of temperatures result in the growth of 共a兲 diamond or 共b兲 amorphous car- were reported to be oriented very differently at the two in-
bon. The molecular structure shows the hydrogen terminated 共1⫻1兲 bulk
structure whose C–H stretching vibration is detected at 2820 cm⫺1 共after
terfaces, more upright at the air/PS interface and lying flat at
Ref. 130兲. the sapphire/PS interface, respectively.44 This experiment is
also interesting from two other points of view. It is one of the
very few cases in which the range of aromatic vibrations has
cm⫺1 关Fig. 9共a兲兴.130 This single peak has already been ob- been investigated. The other point was discussed in Sec. III B
served under UHV conditions131,132 and represents the C–H and relates to the geometry of the setup which allowed se-
stretching vibrations of the hydrogen terminated 共1⫻1兲 sur- lective study of both interfaces of the film. Another example
face. In contrast, under conditions which are unfavorable for is illustrated in Fig. 10 which shows a poly共ethyleneglycol兲
diamond growth 关Fig. 9共b兲兴 this intensity is significantly de- 共PEG兲 layer on CaF2 . 42 This material, which is of interest for
creased and indicates a major change of the surface structure. medical applications,137 was investigated as a function of its
This is corroborated by the emergence of a second band environment. In air 关Fig. 10共a兲兴 a variety of peaks can clearly
around 2855 cm⫺1 which has also been observed under UHV be identified. Besides the symmetric and asymmetric meth-
conditions131,133 and is indicative of a distorted surface struc- ylene vibrations the peaks reflect the presence of different
ture whose details have yet to be unraveled. conformations. Changing the environment from air to water
As pointed out in Sec. IV SFG reveals its unique features results in striking changes in the SFG spectrum 关Fig. 10共b兲兴.
if surfaces of liquids and organic materials such as polymers The water interacts strongly with the PEG and gives rise to
are probed. It is worth mentioning that SFG allowed the first significant structural changes. It is worth mentioning that in
observation of a vibrational spectrum of a neat liquid this experiment the OH stretching region was also probed
surface134 and since then it has provided revealing insights and the spectra indicate significant coordination of the water
into the structure and properties of this type of surface. Liq- molecules at the interface. This example provides evidence
uid surfaces, in particular water, and liquid/liquid interfaces that in situ measurements are mandatory. At the same time it

J. Vac. Sci. Technol. A, Vol. 19, No. 6, NovÕDec 2001

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2731 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2731

also points to an issue which can be generalized for SFG. In


order to make full use of the rich information contained in
the vibrational spectra of such complicated systems, theoret-
ical investigations have to parallel the experiments. This con-
cerns both precise assignment of vibrational bands and mod-
eling of the structure of these systems at different levels of
theory. For ethylene glycol terminated monolayers such a
comprehensive approach was undertaken and has revealed
understanding on a qualitative basis from the perspective of a
more quantitative description.41,138 –140
The previous example showed SFG as a probe of buried
interfaces. A similar situation is encountered in tribological
studies where films are confined between two solids. The
boundary lubrication of ultrathin organic films is of great
interest and methods are needed to study in situ the behavior
of molecules subject to normal and shear forces. Again, the
interface sensitivity of second order optical techniques offers
a unique opportunity by which to study the reorientation and
alignment of confined layers. The sensitivity of studying
them by SFG has been demonstrated141,142 and SHG experi-
ments which can be extended to SFG have pinpointed the
effect of forces on the orientation and lateral alignment of
molecules in point contact.143,144 Here, in particular, the ad-
vantage of focusing down to small spots comes into play
since point contacts associated with high pressures and
forces are more important than large uniform areas. In the
field of tribology there is ample room for further SFG ex-
FIG. 11. SFG microscope image of a monolayer of erythro-rac-9,10-
periments to study boundary lubrication and the behavior of
dihydroxyoctadecanoate on fused silica. The image was recorded with the
coatings on the molecular scale. polarization of the visible at 532 nm and the IR beams parallel to the surface
Whereas the examples presented so far focus on the static plane. The IR frequency was set to 2962 cm⫺1 to probe the asymmetric CH3
properties of a system and the macroscopic areas probed, vibration. The overall gradient in the average SF intensity towards the edge
of the image is due to inhomogeneous illumination of the area. The sharp
extension to both spatial and temporal resolution opens up contrast between bright and dark areas which corresponds to a difference in
additional promising fields of application for SFG. At inter- SF intensity of 21% reflects the presence of two different phases with dif-
faces one often faces the problem that the surface composi- ferent densities and different molecular orientations. 共Figure adapted from
tion is inhomogeneous. For example, in Langmuir–Blodgett Ref. 145.兲
layers different phases can coexist or, as revealed by peeling
tests, the adhesion of polymers can be somewhat inhomoge- unique opportunities for studies of dynamics and kinetics at
neous. Furthermore, in the study of arrays of clusters, where, interfaces and fs pulse technology have opened up new di-
depending on the technique applied, the preparation of mac- mensions. The time scales spanned can vary over several
roscopically large areas can be difficult, probing of small orders of magnitude depending on the coupling to decay
areas might be crucial. Whereas the diffraction limit of nor- channels. For strong coupling, as is found for adsorbates on
mal IR spectroscopy allows, at best, 5–10 ␮m resolution, the metals, the relaxation times are in the range of ⬃1 ps;9 for
upconversion of SFG into the visible range improves this by weak coupling such as hydrogen on Si共111兲,149 it can reach
an order of magnitude and, therefore, its spatial resolution is the nanosecond range. So far rather few reports have been
comparable to that of Raman microscopy. The feasibility of published but with the commercial availability of laser sys-
nonlinear optical microscopy has mainly been demonstrated tems generating ultrashort pulses 共⬍200 fs兲 this is expected
so far with second harmonic generation145,146 but it is also to change. A very interesting extension of time resolved
applicable to SFG as Fig. 11 illustrates.147 A laser system spectroscopy is not only to study pure energy transfer but to
using ns pulses was used to probe the structure of an organic trace chemical changes and, thus, to elucidate individual
monolayer on a solid substrate. Even though the spatial reso- steps in chemical reactions. The study of formic acid on Ni
lution was well below what can be achieved theoretically it and NiO surfaces illustrates the possibility to trace interme-
proves the concept and it is expected that further improve- diate species.150–152 By producing a transient temperature
ment of the experimental geometry will push the resolution jump by heating the substrate with a ps laser pulse, reversible
to significantly smaller dimensions. Efforts to extend the spa- transformation of the formate species from a bidentate to a
tial resolution even to subwavelength dimensions by combin- monodentate and the decomposition of the formate were
ing SFG with near-field optics are underway.148 monitored by SFG. Another very recent example in which
With respect to the temporal resolution SFG offers, femtosecond pulses were applied is depicted in Fig. 12

JVST A - Vacuum, Surfaces, and Films

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2732 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2732

interfacial chemical reactions and to probe the dynamics at


interfaces.

VI. FUTURE DIRECTIONS AND CHALLENGES


A decade of using SFG has provided enlightening insights
into a number of processes on non-UHV surfaces that were
previously inaccessible at this level of detail and the question
now is in what areas will SFG be likely to continue to con-
tribute substantially. Many experiments so far have been car-
ried out on rather simple model systems. Since it can be
expected that use of SFG continues to require less expertise
in laser technology and, importantly, that the laser systems
further improve in reliability, more complicated systems can
be tackled. Among them heterogenous catalysis will likely to
be a major field in which the direction has been
indicated.75,129 A wealth of information can also be expected
for polymers to solve problems in the understanding of ad-
FIG. 12. Transient SFG spectra of CO/Ru共100兲 during desorption. The de- ditives and interfacial segregation phenomena. The tribologi-
sorption is initiated by a fs pulse at 800 nm and the SFG spectra are moni- cal properties of organic layers are also a field in which use
tored as a function of the delay time between the pump and the IR probe of SFG can be expected to make a significant contribution to
pulse. For details see the text. 共Figure adapted from Ref. 153.兲 understanding of the relationship between macroscopic
forces and processes on the molecular scale.
Furthermore, studies of the dynamics at interfaces will
provide a deeper insight into the chemistry at surfaces which,
which shows transient SFG spectra of the intramolecular again, is of utmost importance for catalysis. All experiments
stretching mode of CO on Ru共100兲.53,153 A 110 fs pump pulse so far were performed on flat surfaces but this need not nec-
at 800 nm excites the substrate electrons followed by their essarily have to be the case.154 –156 Another perspective is the
fast equilibration with the substrate phonons which couple detection of chirality at surfaces. Having been demonstrated
efficiently to the CO and cause the molecules to desorb. The with SHG, a SFG result is still to be reported. Only from
SFG signal at 690 nm was generated by a ps visible 共800 bulk chiral liquids was it unequivocally very recently
nm兲 pulse of 4 cm⫺1 spectral width and a fs broad-band IR demonstrated.157
pulse. The series of spectra show the CO signal as a function The strength of SFG at the moment definitely lies in its
of the delay time between the intensive pump pulse and the capability to deduce qualitative information from the assign-
weak SFG probe. As can be seen clearly the vibrational band ment and positions of vibrational bands and its use will con-
exhibits transient changes in intensity, band width, and posi- tinue to make significant contributions at this level. How-
tion. Note that the onset of spectral shifts at negative delay ever, quantification of the signal remains a major challenge.
times is due to perturbations in the free induction decay. As The difficulties encountered in the quantitative evaluation of
reasoned in detail53,153 the frequency shifts reflect the trans- SFG spectra was addressed by Bell et al.33 who analyzed the
fer of energy from the substrate to the CO molecule which molecular orientation based on different C–H stretching
finally causes a significant fraction of the CO molecules to modes. Other attempts have been made that focus on the
desorb as inferred by the decrease of the SFG between the orientation of methyl or methylene groups, mostly in Lang-
first pulse at ⫺4.5 ps and that at 163 ps. muir and thiol films25,26,34 –36,39 but also in polymers such as
In contrast to the last example, in which the visible probe polyvinylalcohol.158 Progress in the quantification can only
pulse is spectrally narrow, another interesting approach com- be expected if theory can reliably calculate the relevant
bined short pulses of both the IR and the visible light.54 The quantities. This will include all levels of modeling from ab
broad band SFG signal was detected as a function of the initio calculations to molecular dynamics. This probably is at
delay between the IR pulse, which excited the vibrations of a the moment the bottleneck since the interpretation of ␹ re-
Langmuir–Blodgett layer of ferric stearate over the range of quires ␣ and ␮ to be known. Already being a formidable task
the C–H stretching vibrations, and the broad band visible for an isolated molecule things worsen for adsorbed mol-
pulse. The signal decay which provides information on vi- ecules due to both interaction with the substrate and intermo-
brational dephasing exhibits pronounced oscillations due to lecular interactions. For larger organic molecules with many
quantum interference of coherently excited vibrations. The conformational degrees of freedom molecular modeling of a
signal decay depends on the kind of treatment of the sample layer may be required to interpret the spectra. Here, a par-
and, thus, the experiment shows a new way of probing the ticular aspect is the argument of inversion symmetry. Quali-
local environment of adsorbates. These examples clearly tatively it is easy to differentiate between vanishing and non-
demonstrate that use of SFG can be expected to make further vanishing signals that indicate centrosymmetric or
significant contributions to trace short lived intermediates in anisotropic structures, respectively. But how is this inter-

J. Vac. Sci. Technol. A, Vol. 19, No. 6, NovÕDec 2001

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2733 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2733

23
preted quantitatively, e.g., for hydrocarbon chains where the s and p denote the polarization parallel to the surface and parallel to the
plane of incidence, respectively. Ordered by decreasing frequency the
signal from the methylene groups is a marker for order/
combination ssp defines an s polarized SFG signal, an s-polarized beam
disorder? Initial attempts to support the interpretation of SFG at ␻ 1 , and a p-polarized beam at ␻ 2 ⬍ ␻ 1 .
data by theory have been made41,138 –140,159 but there is still a 24
D. Zhang, J. Gutov, and K. B. Eisenthal, J. Phys. Chem. 98, 13729
significant step that must be made towards quantitative inter- 共1994兲.
25
C. Hirose, N. Akamatsu, and K. Domen, J. Chem. Phys. 96, 997 共1992兲.
pretation of a SFG signal. 26
C. Hirose, H. Yamamoto, N. Akamatsu, and K. Domen, J. Phys. Chem.
Plenty of room for further development is concerned with 97, 10064 共1993兲.
extension of the wavelength range. Here, an unresolved issue 27
P. Ye and Y. R. Shen, Phys. Rev. B 28, 4288 共1983兲.
is that of how much SFG can contribute to studies of the
28
D. Roy, Phys. Rev. B 61, 13283 共2000兲.
29
Y. R. Shen, Appl. Phys. B: Lasers Opt. B68, 295 共1999兲.
interaction of adsorbates with a substrate which is at the low 30
P. Guyot-Sionnest, J. H. Hunt, and Y. R. Shen, Phys. Rev. Lett. 59, 1597
frequency edge of what is currently achieved with free elec- 共1987兲.
tron lasers. As mentioned earlier, detecting adsorbate– 31
T. H. Ong, P. B. Davies, and C. D. Bain, Langmuir 9, 1836 共1993兲.
32
substrate vibrations has not yet been successful.66 However, M. G. Brown, E. A. Raymond, H. C. Allen, L. F. Scatena, and G. L.
Richmond, J. Phys. Chem. A 104, 10220 共2000兲.
there is a wide range left from about 1700 to about 500 cm⫺1 33
G. R. Bell, Z. X. Li, C. D. Bain, R. K. Thomas, and D. C. Duffy, J. Phys.
which is essentially unexplored but is of great interest since Chem. B 102, 9461 共1998兲.
34
most of the chemical group frequencies appear in his range. A. Lampert, Ph.D. thesis, Ruprecht Karls Universität Heidelberg, Heidel-
The range down to 1000 cm⫺1 is already readily available 35
berg, 1997.
C. Hirose, N. Akamatsu, and K. Domen, Appl. Spectrosc. 46, 1051
with benchtop lasers, leaving plenty of room for studies in 共1992兲.
which SFG will be a key technique by which to study chem- 36
G. R. Bell, C. D. Bain, and R. N. Ward, J. Chem. Soc., Faraday Trans. 92,
istry and physics at interfaces. 515 共1996兲.
37
A. Goto, N. Akamatsu, K. Domen, and C. Hirose, J. Phys. Chem. 99,
4086 共1995兲.
38
J. Löbau and K. Wolfum, J. Opt. Soc. Am. B 14, 2505 共1997兲.
39
ACKNOWLEDGMENTS X. Zhuang, P. B. Miranda, D. Kim, and Y. R. Shen, Phys. Rev. B 59,
12632 共1999兲.
40
The authors are grateful to M. Grunze for his continuous M. Himmelhaus, F. Eisert, M. Buck, and M. Grunze, J. Phys. Chem. B
support, to E. Kay for stimulating the writing of this article, 104, 576 共2000兲.
41
M. Zolk, F. Eisert, J. Pipper, S. Herrwerth, W. Eck, M. Buck, and M.
and to V. Ebert, U. Metka, H. R. Volpp, and C. D. Bain for Grunze, Langmuir 16, 5849 共2000兲.
helpful discussions. Partial support by the German Science 42
L. Dreesen, C. Humbert, P. Hollander, A. A. Mani, K. Ataka, P. A. Thiry,
Foundation is gratefully acknowledged. and A. Peremans, Chem. Phys. Lett. 333, 327 共2001兲.
43
D. Zhang, R. S. Ward, Y. R. Shen, and G. Somorjai, J. Phys. Chem. B
101, 9060 共1997兲.
44
K. S. Gautam, A. D. Schwab, A. Dhinojwala, D. Zhang, S. M. Dougal,
1
X. D. Zhu, H. Suhr, and Y. R. Shen, Phys. Rev. B 35, 3047 共1987兲. and M. S. Yeganeh, Phys. Rev. Lett. 85, 3854 共2000兲.
2
E. Matthias and F. Träger, Eds., Appl. Phys. B 68共3兲 共1999兲. 45
P. B. Miranda, L. Xu, Y. R. Shen, and M. Salmeron, Phys. Rev. Lett. 81,
3
C. D. Bain, J. Chem. Soc., Faraday Trans. 91, 1281 共1995兲. 5876 共1998兲.
4 46
C. D. Bain, in Modern Characterization Methods of Surfactant Systems, X. C. Su, L. Lianos, Y. R. Shen, and G. A. Somorjai, Phys. Rev. Lett. 80,
edited by B. P. Binks 共Dekker, New York, 1999兲, p. 336. 1533 共1998兲.
5
R. B. Hall, J. N. Russell, J. Miragliotta, and P. R. Rabinowitz, in Chem- 47
O. Dannenberger, J. J. Wolff, and M. Buck, Langmuir 14, 4679 共1998兲.
48
istry and Physics of Solid Surfaces VIII, edited by R. Vanselow and R. M. Buck, F. Eisert, J. Fischer, M. Grunze, and F. Träger, J. Vac. Sci.
Howe 共Springer, Berlin, 1991兲, Vol. 22, p. 87. Technol. A 10, 926 共1992兲.
6 49
A. Tadjeddine and A. Peremans, Spectroscopy for Surface Science, edited M. Buck, F. Eisert, M. Grunze, and F. Träger, Appl. Phys. A: Mater. Sci.
by R. J. H. Clark and R. E. Hester 共Wiley, Chichester, 1998兲, Vol. 26, p. Process. A60, 1 共1995兲.
159. 50
A. Liebsch, Appl. Phys. B 68, 301 共1999兲.
7
P. B. Miranda and Y. R. Shen, J. Phys. Chem. B 103, 3292 共1999兲. 51
E. W. M. van der Ham, Q. H. F. Vrehen, and E. R. Eliel, Opt. Lett. 21,
8
D. E. Gragson and G. L. Richmond, J. Phys. Chem. B 102, 3847 共1998兲. 1448 共1996兲.
9 52
H. L. Dai and W. Ho, Surface Vibrational Dynamics Probed by Sum L. J. Richter, T. P. Petralli-Mallow, and J. C. Stephenson, Opt. Lett. 23,
Frequency Generation, Laser Spectroscopy and Photochemistry on Metal 1594 共1998兲.
Surfaces Part 1 共1995兲. 53
C. Hess, S. Funk, M. Bonn, D. N. Denzler, M. Wolf, and G. Ertl, Appl.
10
C. Hirose, Adv. Multiphoton Process. Spectrosc. 9, 145 共1995兲. Phys. A: Mater. Sci. Process. A71, 477 共2000兲.
11
M. J. Shultz, C. Schnitzer, D. Simonelli, and S. Baldelli, Int. Rev. Phys. 54
D. Star, T. Kikteva, and G. W. Leach, J. Chem. Phys. 111, 14 共1999兲.
Chem. 19, 123 共2000兲. 55
C. D. Bain, P. B. Davies, T. H. Ong, and R. N. Ward, Langmuir 7, 1563
12
P. Dumas, M. K. Weldon, Y. J. Chabal, and G. P. Williams, Surf. Rev. 共1991兲.
Lett. 6, 225 共1999兲. 56
P. Brechignac, S. De Benedictis, H. Nguyen Dai, and N. Halberstadt, Rev.
13
V. Vogel, Curr. Opin. Colloid Interface Sci. 1, 257 共1996兲. Phys. Appl. 21, 735 共1986兲.
14
Y. R. Shen, Solid State Commun. 102, 221 共1997兲. 57
S. J. Brosnan, R. N. Fleming, R. L. Herbst, and R. L. Byer, Appl. Phys.
15
G. Somorjai and G. Rupprechter, J. Phys. Chem. B 103, 1623 共1999兲. Lett. 30, 330 共1977兲.
16
A. Tadjeddine and A. Peremans, Surf. Sci. 368, 377 共1996兲. 58
H. J. Krause and W. Daum, Appl. Phys. B: Photophys. Laser Chem. B56,
17
N. Bloembergen, Nonlinear Optics 共Addison–Wesley, Redwood City, 8 共1993兲.
59
CA, 1992兲. D. E. Gragson, B. M. Mc Carty, G. L. Richmond, and D. S. Alavi, J. Opt.
18
Y. R. Shen, The Principles of Nonlinear Optics 共Wiley, New York, 1984兲. Soc. Am. B 13, 2075 共1996兲.
19
P. N. Butcher and D. Cotter, The Elements of Nonlinear Optics 共Cam- 60
A. Seilmeier, K. Spanner, A. Laubereau, and W. Kaiser, Opt. Commun.
bridge University Press, Cambridge, 1990兲. 24, 237 共1978兲.
20
D. L. Mills, Nonlinear Optics: Basic Concepts 共Springer, Berlin, 1998兲. 61
D. C. Hanna, V. V. Rampal, and R. C. Smith, Opt. Commun. 8, 151
21
Y. R. Shen, Annu. Rev. Phys. Chem. 40, 327 共1989兲. 共1973兲.
22
P. F. Brevet, Surface Second Harmonic Generation 共Polytechniques Uni- 62
Y. Caudano, A. Peremans, P. A. Thiry, P. Dumas, and A. Tadjeddine, Surf.
versitaires Romandes, Lausanne, 1997兲. Sci. 368, 337 共1996兲.

JVST A - Vacuum, Surfaces, and Films

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2734 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2734

63
A. Peremans, A. Tadjeddine, W.-Q. Zheng, A. L. Rille, P. Guyot-Sionnest, 105
Y. Chabal, Surf. Sci. Rep. 8, 211 共1988兲.
and P. A. Thiry, Surf. Sci. 368, 384 共1996兲. 106
P. Dumas, Surf. Interface Anal. 22, 561 共1994兲.
64 107
E. R. Eliel, E. W. M. van der Ham, Q. H. F. Vrehen, G. W. t’Hooft, M. S. Katano, A. Bandara, J. Kubota, K. Onda, A. Wada, K. Domen, and C.
Barmentlo, J. M. Auerhammer, A. F. G. van der Meer, and P. W. van Hirose, Surf. Sci. 428, 337 共1999兲.
Amersfoort, Appl. Phys. A: Mater. Sci. Process. A60, 113 共1995兲. 108
J. A. Creighton, in Spectroscopy of Surfaces, edited by R. J. H. Clarke and
65
S. R. Hatch, R. S. Polizzotti, S. Dougal, and P. Rabinowitz, Chem. Phys. R. E. Hester 共Wiley: Chichester, 1988兲, Vol. 16, p. 37.
Lett. 196, 97 共1992兲. 109
A. Ulman, Chem. Rev. 96, 1533 共1996兲.
66
C. T. Williams, Y. Yang, and C. D. Bain, Catal. Lett. 61, 7 共1999兲. 110
S. R. Goates, D. A. Schofield, and C. D. Bain, Langmuir 15, 1400 共1999兲.
67
A. Dhirani and P. Guyot-Sionnest, Opt. Lett. 20, 1104 共1995兲. 111
M. K. Debe, J. Appl. Phys. 55, 3354 共1984兲.
68
A. A. Mani, P. Hollander, P. A. Thiry, and A. Peremans, Appl. Phys. Lett. 112
A. N. Parikh and D. L. Allara, J. Chem. Phys. 96, 927 共1992兲.
75, 3066 共1999兲. 113
H.-T. Rong, S. Frey, Y.-J. Yang, M. Zharnikov, M. Buck, M. Wuehn, C.
69
U. Keller et al., IEEE J. Sel. Top. Quantum Electron. 2, 435 共1996兲. Woell, and G. Helmchen, Langmuir 17, 1582 共2001兲.
70 114
L. R. Brovelli, I. D. Jung, D. Kopf, M. Kamp, M. Moser, F. X. Kartner, N. Watanabe, H. Yamamoto, A. K. D. Wada, C. Hirose, T. Ohtake, and N.
and U. Keller, Electron. Lett. 31, 287 共1995兲. Mino, Spectrochim. Acta 50, 1529 共1994兲.
71
V. Couderc, O. Guy, E. Roisse, and A. Barthelemy, Electron. Lett. 34, 672 115
K. Wolfrum and A. Laubereau, Chem. Phys. Lett. 228, 83 共1994兲.
共1998兲. 116
G. J. Simpson and K. L. Rowlen, Acc. Chem. Res. 33, 781 共2000兲.
72
D. Negus and C. Seaton, Laser Focus World 28, 69 共1992兲. 117
C. D. Bain, P. B. Davies, and R. N. Ward, Langmuir 10, 2060 共1994兲.
73
G. Cerullo, S. D. Silvestri, and V. Magni, Opt. Lett. 19, 1040 共1994兲. 118
S. Baldelli, G. Mailhot, P. Ross, Y. R. Shen, and G. Somorjai, J. Phys.
74
X. Wei, S.-C. Hong, A. I. Lvovsky, H. Held, and Y. R. Shen, J. Phys. Chem. B 105, 654 共2001兲.
Chem. B 104, 3349 共2000兲. 119
R. N. Ward, P. N. Davies, and C. D. Bain, J. Phys. Chem. 97, 7141
75
T. Dellwig, G. Rupprechter, H. Unterhalt, and H.-J. Freund, Phys. Rev. 共1993兲.
Lett. 85, 776 共2000兲. 120
M. R. Anderson, M. N. Evaniak, and M. Zhang, Langmuir 12, 2327
76
In fact, the efficiency goes down again when going to very short pulses 共1996兲.
121
due to decreasing efficiency of the conversion processes used to generate D. Blaudez, T. Buffeteau, B. Desbat, and J. M. Turlet, Curr. Opin. Colloid
tunable IR. A detailed comparison of fs with ps has to consider the trade- Interface Sci. 4, 265 共1999兲.
122
off between the decreasing conversion efficiency and the time saved for A. Tadjeddine, A. Peremans, A. Le Rille, W.-Q. Zheng, and J.-P. Flament,
tuning the IR wavelength. J. Chem. Soc., Faraday Trans. 92, 3823 共1996兲.
77
W. Koechner, Solid-State Laser Engineering 共Springer, Berlin, 1999兲, 123
Hytran Optical Spectra Database Ver. 2.3, ONTAR Corp, North Andover,
Vol. 1, Chap. 11. MA, 1994.
78 124
G. A. Sefler, P. B. Miranda, and Y. R. Shen, Chem. Phys. Lett. 235, 347 G. Rupprechter, T. Dellwig, H. Unterhalt, and H.-J. Freund, J. Phys.
共1995兲. Chem. B 105, 3797 共2001兲.
79
Y. R. Shen, The Principles of Nonlinear Optics 共Wiley, New York, 1984兲, 125
H. L. Fang and R. L. Swofford, in Ultrasensitive Spectroscopy, edited by
p. 72, Eq. 6.11. D. S. Kliger 共Academic, New York, 1983兲, p. 176.
80
E. A. Potterton and C. D. Bain, J. Electroanal. Chem. 409, 109 共1996兲. 126
P. S. Cremer, X. C. Su, Y. R. Shen, and G. A. Somorjai, J. Am. Chem.
81
B. Dick, A. Gierulski, G. Marowsky, and G. A. Reider, Appl. Phys. B 38, Soc. 118, 2942 共1996兲.
107 共1985兲. 127
X. C. Su, P. S. Cremer, Y. R. Shen, and G. A. Somorjai, J. Am. Chem.
82
D. A. Beattie, S. Haydock, and C. D. Bain, Vib. Spectrosc. 24, 109 Soc. 119, 3994 共1997兲.
共2000兲. 128
H. Harle, U. Metka, H. R. Volpp, and J. Wolfrum, Phys. Chem. Chem.
83
M. C. Messmer, J. C. Conboy, and G. L. Richmond, J. Am. Chem. Soc. Phys. 1, 5059 共1999兲.
117, 8039 共1995兲. 129
S. Baldelli, A. S. Eppler, E. Anderson, Y. R. Shen, and G. A. Somorjai, J.
84
J. C. Conboy, M. C. Messmer, and G. L. Richmond, J. Phys. Chem. 100, Chem. Phys. 113, 5432 共2000兲.
7617 共1996兲. 130
A. Heerwagen, M. Strobel, M. Himmelhaus, and M. Buck, J. Am. Chem.
85
C. T. Williams, Y. Yang, and C. D. Bain, Langmuir 16, 2343 共2000兲. Soc. 123, 6732 共2001兲.
86
H. Raether, Surface Plasmons 共Springer, Berlin, 1988兲. 131
R. P. Chin, J. Y. Huang, R. Y. Shen, T. J. Chuang, H. Seki, and M. Buck,
87
E. V. Alieva, L. A. Kuzik, V. A. Yakovlev, G. Knippels, A. F. G. van der Phys. Rev. B 45, 1522 共1992兲.
Meer, and G. Mattei, Chem. Phys. Lett. 302, 528 共1999兲. 132
R. P. Chin, J. Y. Huang, Y. R. Shen, T. J. Chuang, and H. Seki, Phys. Rev.
88
E. R. Eliel, E. W. M. van der Ham, and Q. H. F. Vrehen, Appl. Phys. B: B 52, 5985 共1995兲.
Lasers Opt. 68, 349 共1999兲. 133
R. P. Chin, J. Y. Huang, Y. R. Shen, T. J. Chuang, and H. Seki, Phys. Rev.
89
E. W. M. van der Ham, Q. H. E. Vrehen, E. R. Eliel, V. A. Yakovlev, E. V. B 54, 8243 共1996兲.
134
Valieva, L. A. Kuzik, J. E. Petrov, V. A. Sychugov, and A. F. G. van der R. Superfine, J. Y. Huang, and Y. R. Shen, Phys. Rev. Lett. 66, 1066
Meer, J. Opt. Soc. Am. B 16, 1146 共1999兲. 共1991兲.
90
M. A. Hines, J. A. Todd, and P. Guyot-Sionnest, Langmuir 11, 493 共1995兲. 135
J. C. Conboy, M. C. Messmer, and G. L. Richmond, Langmuir 14, 6722
91
Y. R. Shen, in Ref. 79, Sec. 6. 共1998兲.
92
A. Tadjeddine, A. Le Rille, O. Pluchery, F. Vidal, W. Q. Zheng, and A. 136
M. Buck, C. Dressler, M. Grunze, and F. Träger, J. Adhes. 58, 227 共1996兲.
Peremans, Phys. Status Solidi A 175, 89 共1999兲. 137
J. M. Harris, Poly(ethyleneglycol)chemistry: Biotechnical and Biomedical
93
A. Tadjeddine and A. L. Rille, Electrochim. Acta 45, 601 共1999兲. Applications 共Plenum, New York, 1992兲.
94
A. L. Rille and A. Tadjeddine, J. Electroanal. Chem. 467, 238 共1999兲. 138
R. L. C. Wang, H. J. Kreuzer, and M. Grunze, Phys. Chem. Chem. Phys.
95
W. L. Mochan, J. A. Maytorena, and B. S. Mendoza, Phys. Status Solidi 2, 3613 共2000兲.
A 170, 357 共1998兲. 139
A. Pertsin and M. Grunze, Langmuir 16, 8829 共2000兲.
96
B. S. Mendoza, W. L. Mochan, and J. A. Maytorena, Phys. Rev. B 60, 140
M. Buck 共submitted兲.
14334 共1999兲. 141
Q. Du, X. D. Xiao, D. Charych, F. Wolf, P. Frantz, Y. R. Shen, and M.
97
J. Y. Huang and Y. R. Shen, Phys. Rev. A 49, 3973 共1994兲. Salmeron, Phys. Rev. B 51, 7456 共1995兲.
98
M. Buck and A. Peremans 共unpublished兲. 142
R. Fraenkel, G. E. Butterworth, and C. D. Bain, J. Am. Chem. Soc. 120,
99
A. Otto, J. Phys.: Condens. Matter 4, 1143 共1992兲. 203 共1998兲.
100 143
M. A. Hines, T. D. Harris, and Y. J. Chabal, J. Electron Spectrosc. Relat. F. Eisert, M. Gurka, A. Legant, M. Buck, and M. Grunze, Science 287,
Phenom. 64Õ65, 183 共1993兲. 468 共2000兲.
101
H. Sano and S. Ushioda, Phys. Rev. B 53, 1958 共1996兲. 144
M. Strobel, M. Buck, J. Blümmel, W. Eck, and M. Grunze, 24th Annual
102
M. Cai, M. D. Mowery, J. E. Pemberton, and C. E. Evans, Appl. Spec- Meeting of the Adhesion Society, 25–28 February 2001, Williamsburg,
trosc. 54, 31 共2000兲. VA, 2001.
103
A. Nemetz, T. Fischer, A. Ulman, and W. Knoll, J. Chem. Phys. 98, 5912 145
M. Floersheimer, Phys. Status Solidi A 173, 15 共1999兲.
共1993兲. 146
C. E. Allen, R. Ditchfield, and E. G. Seebauer, Phys. Rev. B 55, 13304
104
T. Kawai, J. Unemura, and T. Takenaka, Chem. Phys. Lett. 162, 243 共1997兲.
共1989兲. 147
M. Floersheimer, C. Brillert, and H. Fuchs, Langmuir 15, 5437 共1999兲.

J. Vac. Sci. Technol. A, Vol. 19, No. 6, NovÕDec 2001

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2735 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2735

148 188
B. Humbert, J. Grausem, A. Burneau, M. Spajer, and A. Tadjeddine, Appl. D. H. Gracias, D. Zhang, L. Lianos, W. Ibach, Y. R. Shen, and G. A.
Phys. Lett. 78, 135 共2001兲. Somorjai, Chem. Phys. 245, 277 共1999兲.
149 189
P. Guyot-Sionnest, P. Dumas, and Y. J. Chabal, J. Electron Spectrosc. Z. Chen, R. Ward, Y. Tian, A. S. Eppler, Y. R. Shen, and G. A. Somorjai,
Relat. Phenom. 54, 27 共1990兲. J. Phys. Chem. B 103, 2935 共1999兲.
150 190
C. Hirose, A. Bandara, S. Katano, J. Kubota, A. Wada, and K. Domen, D. Zhang, D. H. Gracias, R. Ward, M. Gauckler, Y. Tian, Y. R. Shen, and
Appl. Phys. B: Lasers Opt. 68, 559 共1999兲. G. A. Somorjai, J. Phys. Chem. B 102, 6225 共1998兲.
151
A. Bandara, J. Kubota, K. Onda, A. Wada, K. Domen, and C. Hirose, 191
D. Kim and Y. R. Shen, Appl. Phys. Lett. 74, 3314 共1999兲.
Surf. Sci. 435, 83 共1999兲. 192
D. Zhang, S. M. Dougal, and M. S. Yeganeh, Langmuir 16, 4528 共2000兲.
152
A. Bandara, J. Kubota, A. Wada, K. Domen, and C. Hirose, Appl. Phys. 193
M. Himmelhaus, M. Buck, and M. Grunze, Appl. Phys. B 68, 595 共1999兲.
B: Lasers Opt. B68, 573 共1999兲. 194
Y. Tanaka, S. Lin, N. Aono, and T. Suzuki, Appl. Phys. B: Lasers Opt. 68,
153
M. Bonn, C. Hess, J. H. Miners, B. N. J. Persson, M. Wolf, and G. Ertl, 713 共1999兲.
Phys. Rev. Lett. 84, 4653 共2000兲. 195
R. Braun, B. D. Casson, C. D. Bain, E. W. M. van der Ham, Q. H. F.
154
H. Wang, T. Troxler, A.-G. Yeh, and H.-L. Dai, Langmuir 16, 2475 Vrehen, E. R. Eliel, A. M. Briggs, and P. B. Davies, J. Chem. Phys. 110,
共2000兲. 4634 共1999兲.
155 196
H. F. Wang, E. C. Y. Yan, Y. Liu, and K. B. Eisenthal, J. Phys. Chem. 102, A. L. Harris, L. Rothberg, L. Dhar, N. J. Levinos, and L. H. Dubois, J.
4446 共1998兲. Chem. Phys. 94, 2438 共1991兲.
156
H. F. Wang, E. C. Y. Yan, E. Borguet, and K. B. Eisenthal, Chem. Phys. 197
M. S. Yeganeh, S. M. Dougal, R. S. Polizzotti, and P. Rabinowitz, Thin
Lett. 259, 15 共1996兲. Solid Films 270, 226 共1995兲.
157
M. A. Belkin, T. A. Kulakov, K.-H. Ernst, L. Yan, and Y. R. Shen, Phys. 198
I. Böhm, A. Lampert, M. Buck, F. Eisert, and M. Grunze, Appl. Surf. Sci.
Rev. Lett. 85, 4474 共2000兲. 141, 237 共1999兲.
158
X. Wei, X. W. Zhuang, S. C. Hong, T. Goto, and Y. R. Shen, Phys. Rev. 199
Y.-S. Shon, R. Colorado, C. T. Williams, C. D. Bain, and T. R. Lee,
Lett. 82, 4256 共1999兲. Langmuir 16, 541 共2000兲.
159
A. Morita and J. T. Hynes, Chem. Phys. 258, 371 共2000兲. 200
R. Superfine, P. Guyot-Sionnest, J. H. Hunt, C. T. Kao, and Y. R. Shen,
160
Q. Du, R. Superfine, E. Freysz, and Y. R. Shen, Phys. Rev. Lett. 70, 2313 Surf. Sci. 200, L445 共1988兲.
共1993兲. 201
P. Guyot-Sionnest, R. Superfine, J. H. Hunt, and Y. R. Shen, Chem. Phys.
161
D. E. Gragson and G. L. Richmond, Langmuir 13, 4804 共1997兲. Lett. 144, 1 共1988兲.
162
Q. Du, E. Freysz, and Y. R. Shen, Science 264, 826 共1994兲. 202
J. Y. Huang, K. J. Song, A. Lagoutchev, P. K. Yang, and T. J. Chuang,
163
L. F. Scatena, M. G. Brown, and G. L. Richmond, Science 292, 908 Langmuir 13, 58 共1997兲.
共2001兲. 203
J. Löbau and K. Wolfum, Laser Phys. 8, 582 共1998兲.
164
E. Freysz, Q. Du, and Y. R. Shen, Ann. Phys. 19, 95 共1994兲. 204
M. Sass, J. Loebau, M. Lettenberger, and A. Laubereau, Chem. Phys.
165
R. A. Walker, D. E. Gragson, and G. L. Richmond, Colloids Surf., A 154, Lett. 311, 13 共1999兲.
175 共1999兲. 205
A. L. Harris, C. E. D. Chidsey, N. J. Levinos, and D. N. Loiacono, Chem.
166
D. E. Gragson, B. M. McCarty, and G. L. Richmond, J. Am. Chem. Soc. Phys. Lett. 141, 350 共1987兲.
119, 6144 共1997兲. 206
P. Guyot-Sionnest, J. H. Hunt, and Y. R. Shen, Chem. Phys. Lett. 133, 189
167
D. E. Gragson and G. L. Richmond, J. Am. Chem. Soc. 120, 366 共1998兲. 共1987兲.
168
D. E. Gragson and G. L. Richmond, J. Chem. Phys. 107, 9687 共1997兲. 207
A. L. Harris and N. J. Levinos, J. Chem. Phys. 90, 3878 共1989兲.
169
D. E. Gragson and G. L. Richmond, J. Phys. Chem. B 102, 569 共1998兲. 208
K. Wolfrum, J. Loebau, W. Birkhoelzer, and A. Laubereau, J. Opt. B:
170
C. Raduge, V. Pflumio, and Y. R. Shen, Chem. Phys. Lett. 274, 140 Quantum Semiclassical Opt. 9, 257 共1997兲.
共1997兲.
171
209
M. M. Knock and C. D. Bain, Langmuir 16, 2857 共2000兲.
C. Schnitzer, S. Baldelli, and M. J. Shultz, Chem. Phys. Lett. 313, 416 210
B. D. Casson and C. D. Bain, J. Phys. Chem. B 103, 4678 共1999兲.
共1999兲.
172
211
B. D. Casson and C. D. Bain, J. Phys. Chem. B 102, 7434 共1998兲.
S. Baldelli, D. J. Campbell, C. Schnitzer, and M. J. Shultz, J. Phys. Chem. 212
R. Braun, B. D. Casson, and C. D. Bain, Chem. Phys. Lett. 245, 326
B 103, 2789 共1999兲.
173 共1995兲.
S. Baldelli, D. J. Campbell, C. Schnitzer, and M. J. Shultz, Chem. Phys. 213
B. D. Casson, R. Braun, and C. D. Bain, Faraday Discuss. 104, 209
Lett. 287, 143 共1998兲.
174 共1997兲.
S. Baldelli, C. Schnitzer, M. J. Shultz, and D. J. Campbell, J. Phys. Chem. 214
B. Berge and A. Renault, Europhys. Lett. 21, 773 共1993兲.
B 101, 10435 共1997兲.
175
C. Schnitzer, S. Baldelli, and M. J. Shultz, J. Phys. Chem. B 103, 6383
215
P. B. Miranda, Q. Du, and Y. R. Shen, Chem. Phys. Lett. 286, 1 共1998兲.
216
共1999兲. R. N. Ward, D. C. Duffy, G. R. Bell, and C. D. Bain, Mol. Phys. 88, 269
176
H. C. Allen, E. A. Raymond, and G. L. Richmond, J. Phys. Chem. A. 105, 共1996兲.
217
1649 共2001兲. C. R. Maechling, D. A. V. Kliner, and D. Klenerman, Appl. Spectrosc. 47,
177
S. Baldelli, S. Schnitzer, and M. J. Shultz, Chem. Phys. Lett. 302, 157 167 共1993兲.
218
共1999兲. P. B. Miranda, V. Pflumio, H. Saijo, and Y. R. Shen, Chem. Phys. Lett.
178
C. Schnitzer, S. Baldelli, and M. J. Shultz, J. Chem. Phys. 108, 9817 264, 387 共1997兲.
219
共1998兲. P. B. Miranda, V. Pflumio, H. Saijo, and Y. R. Shen, J. Am. Chem. Soc.
179
C. Schnitzer, S. Baldelli, and M. J. Shultz, J. Phys. Chem. B 104, 585 120, 12092 共1998兲.
220
共2000兲. D. C. Duffy, R. N. Ward, P. B. Davies, and C. D. Bain, J. Am. Chem. Soc.
180
K. Wolfrum, H. Graener, and A. Laubereau, Chem. Phys. Lett. 213, 41 116, 1125 共1994兲.
221
共1993兲. H. C. Allen, E. A. Raymond, and G. L. Richmond, Curr. Opin. Colloid
181
J. Huang and M. H. Wu, Phys. Rev. E 50, 3737 共1994兲. Interface Sci. 5, 74 共2000兲.
182 222
C. D. Stanners, Q. Du, R. P. Chin, P. Cremer, G. A. Somorjai, and Y. R. T. Anzai, H. Maeoka, A. Wada, K. Domen, C. Hirose, T. Ando, and Y.
Shen, Chem. Phys. Lett. 232, 407 共1995兲. Sato, J. Mol. Struct. 352, 455 共1995兲.
183 223
S. Baldelli, C. Schnitzer, M. J. Shultz, and D. J. Campbell, J. Phys. Chem. R. P. Chin, X. Blase, Y. R. Shen, and S. G. Louie, Europhys. Lett. 30, 399
B 101, 4607 共1997兲. 共1995兲.
184 224
D. Zhang, J. H. Gutov, and K. B. Eisenthal, J. Chem. Soc., Faraday Trans. T. Yuzawa, J. Kubota, K. Onda, A. Wada, K. Domen, and C. Hirose, J.
92, 539 共1996兲. Mol. Struct. 413Õ414, 307 共1997兲.
185 225
D. Zhang, J. Gutov, K. B. Eisenthal, and T. F. Heinz, J. Chem. Phys. 98, T. Yuzawa, T. Shioda, J. Kubota, K. Onda, A. Wada, K. Domen, and C.
5099 共1993兲. Hirose, Surf. Sci. 416, L1090 共1998兲.
186 226
D. Simonelli, S. Baldelli, and M. J. Shultz, Chem. Phys. Lett. 298, 400 A. Bandara, J. Kubota, K. Onda, A. Wada, S. S. Kano, K. Domen, and C.
共1999兲. Hirose, J. Phys. Chem. B 102, 5951 共1998兲.
187 227
D. Zhang, Y. R. Shen, and G. Somorjai, Chem. Phys. Lett. 281, 394 H. Ishida, K. Iwatsu, J. Kubota, A. Wada, K. Domen, and C. Hirose, Surf.
共1997兲. Sci. 366, L724 共1996兲.

JVST A - Vacuum, Surfaces, and Films

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30
2736 M. Buck and M. Himmelhaus: Vibrational spectroscopy of interfaces 2736

228 256
H. Ishida, K. Iwatsu, J. Kubota, A. Wada, K. Domen, and C. Hirose, J. A. Bandara, S. Katano, J. Kubota, K. Onda, A. Wada, K. Domen, and C.
Chem. Phys. 108, 5957 共1998兲. Hirose, Chem. Phys. Lett. 290, 261 共1998兲.
229 257
C. Hirose, H. Ishida, K. Iwatsu, N. Watanabe, J. Kubota, A. Wada, and K. A. Bandara, S. Dobashi, J. Kubota, K. Onda, A. Wada, K. Domen, C.
Domen, J. Chem. Phys. 108, 5948 共1998兲. Hirose, and S. S. Kano, Surf. Sci. 387, 312 共1997兲.
230
N. Watanabe, K. Iwatsu, A. Yamakata, T. Ohtani, J. Kubota, J. N. Kondo, 258
B. Bourguignon, S. Carrez, B. Dragnea, and H. Dubost, Surf. Sci. 418,
A. Wada, K. Domen, and C. Hirose, Surf. Sci. 358, 651 共1996兲. 171 共1998兲.
231
H. Yamamoto, N. Watanabe, A. Wada, K. Domen, and C. Hirose, J. 259
S. Carrez, B. Dragnea, W.-Q. Zheng, H. Dubost, and B. Bourguignon,
Chem. Phys. 106, 4734 共1997兲. Surf. Sci. 440, 151 共1999兲.
232
K. Domen, H. Yamamoto, N. Watanabe, and A. Wada, Appl. Phys. A: 260
U. Metka, M. G. Schweitzer, H.-R. Volpp, J. Wolfrum, and J. Warnatz, Z.
Mater. Sci. Process. A60, 131 共1995兲. Phys. Chem. 214, 865 共2000兲.
233
P. S. Cremer, X. C. Su, Y. R. Shen, and G. Somorjai, J. Chem. Soc., 261
P. Guyot-Sionnest, Phys. Rev. Lett. 67, 2323 共1991兲.
Faraday Trans. 92, 4717 共1996兲. 262
234 M. Morin, K. Kuhnke, P. Jakob, Y. J. Chabal, N. J. Levinos, and A. L.
P. S. Cremer, X. Su, Y. R. Shen, and G. A. Somorjai, J. Phys. Chem. 100,
Harris, J. Electron Spectrosc. Relat. Phenom. 64Õ65, 11 共1993兲.
16302 共1996兲.
235
263
P. Guyot-Sionnest, J. Electron Spectrosc. Relat. Phenom. 64Õ65, 1 共1993兲.
X. C. Su, Y. R. Shen, and G. A. Somorjai, Chem. Phys. Lett. 280, 302
共1997兲.
264
P. Guyot-Sionnest, Phys. Rev. Lett. 66, 1489 共1991兲.
265
236
X. C. Su, K. Y. Kung, J. Lahtinen, Y. R. Shen, and G. A. Somorjai, J. Mol. Y. J. Chabal, P. Dumas, P. Guyot-Sionnest, and G. S. Higashi, Surf. Sci.
Catal. A: Chem. 141, 9 共1999兲. 242, 524 共1991兲.
266
237
X. C. Su, K. Kung, J. Lahtinen, R. Y. Shen, and G. A. Somorjai, Catal. M. Y. Mao, P. B. Miranda, D. S. Kim, and Y. R. Shen, Appl. Phys. Lett.
Lett. 54, 9 共1998兲. 75, 3357 共1999兲.
267
238
P. Cremer, C. Stanners, J. W. Niemantsverdriet, Y. R. Shen, and G. So- P. Guyot-Sionnest, P. Lin, and E. M. Hiller, J. Chem. Phys. 102, 4269
morjai, Surf. Sci. 328, 111 共1995兲. 共1995兲.
239
P. S. Cremer, X. C. Su, Y. R. Shen, and G. A. Somorjai, J. Phys. Chem. B
268
A. Peremans and A. Tadjeddine, J. Chem. Phys. 103, 7197 共1995兲.
269
101, 6474 共1997兲. A. Peremans, A. Tadjeddine, and P. Guyot-Sionnest, Surf. Sci. 335, 210
240
H. Harle, A. Lehnert, U. Metka, H. R. Volpp, L. Willms, and J. Wolfrum, 共1995兲.
Appl. Phys. B: Lasers Opt. B68, 567 共1999兲. 270
A. Tadjeddine and A. Peremans, J. Electroanal. Chem. 409, 115 共1996兲.
241
H. Harle, A. Lehnert, U. Metka, H.-R. Volpp, L. Willms, and J. Wolfum, 271
W. Daum, F. Dederichs, and J. E. Müller, Phys. Rev. Lett. 80, 766 共1998兲.
Chem. Phys. Lett. 293, 26 共1998兲. 272
K. A. Friedrich, W. Daum, C. Klunker, D. Knabben, U. Stimming, and H.
242
C. Kluenker, M. Balden, S. Lehwald, and W. Daum, Surf. Sci. 360, 104 Ibach, Surf. Sci. 335, 315 共1995兲.
共1996兲. 273
W. Daum, K. A. Friedrich, C. Klunker, D. Knabben, U. Stimming, and H.
243
F. Dederichs, K. A. Friedrich, and W. Daum, J. Phys. Chem. B 104, 6626 Ibach, Appl. Phys. A 59, 553 共1994兲.
共2000兲. 274
W. Daum, F. Dederichs, and J. E. Müller, Phys. Rev. Lett. 85, 2655
244
S. Baldelli, N. Markovic, P. Ross, Y. R. Shen, and G. Somorjai, J. Phys. 共2000兲.
Chem. B 103, 8920 共1999兲. 275
C. Matranga and P. Guyot-Sionnest, J. Chem. Phys. 112, 7615 共2000兲.
245
A. Peremans, A. Tadjeddine, P. Guyot-Sionnest, R. Prazeres, F. Glotin, D. 276
A. Le Rille and A. Tadjeddine, J. Electroanal. Chem. 467, 238 共1999兲.
Jaroszynski, J. M. Berset, and J. M. Ortega, Nucl. Instrum. Methods Phys. 277
A. Le Rille, A. Tadjeddine, W. Q. Zheng, and A. Peremans, Chem. Phys.
Res. A 341, 146 共1994兲.
246 Lett. 271, 95 共1997兲.
A. Tadjeddine, A. Peremans, and P. Guyot-Sionnest, Surf. Sci. 335, 210 278
A. Tadjeddine and P. Guyot-Sionnest, Electrochim. Acta 36, 1849 共1991兲.
共1995兲. 279
O. Pluchery, W. Q. Zheng, T. Marin, and A. Tadjeddine, Phys. Status
247
M. E. Schmidt and P. Guyot-Sionnest, J. Chem. Phys. 104, 2438 共1996兲.
248
H. Harle, K. Mendel, U. Metka, H. R. Volpp, L. Willms, and J. Wolfrum, Solidi A 175, 145 共1999兲.
280
Chem. Phys. Lett. 279, 275 共1997兲. G. A. Browmaker, J.-M. Leger, A. Le Rille, C. A. Melendres, and A.
249
X. C. Su, P. S. Cremer, Y. R. Shen, and G. A. Somorjai, Phys. Rev. Lett. Tadjeddine, J. Chem. Soc., Faraday Trans. 94, 1309 共1998兲.
281
77, 3858 共1996兲. Y. Caudano, A. Peremans, P. A. Thiry, P. Dumas, and A. Tadjeddine, J.
250
A. Peremans, A. Tadjeddine, and P. Guyot-Sionnest, Chem. Phys. Lett. Phys. B 29, 5023 共1996兲.
282
247, 243 共1995兲. A. Peremans, Y. Caudano, P. A. Thiry, P. Dumas, W. Q. Zhang, A. Le
251
P. Guyot-Sionnest and A. Tadjeddine, Chem. Phys. Lett. 172, 341 共1990兲. Rille, and A. Tadjeddine, Phys. Rev. Lett. 78, 2999 共1997兲.
283
252
M. Morin, N. J. Levinos, and A. L. Harris, J. Chem. Phys. 96, 3950 Y. Caudano, A. Peremans, P. A. Thiry, P. Dumas, W. Q. Zheng, A. Le
共1992兲. Rille, and A. Tadjeddine, Phys. Mag. 20, 31 共1998兲.
253
U. Schroder and P. Guyot-Sionnest, Surf. Sci. 421, 53 共1999兲.
284
C. Humbert et al., Phys. Status Solidi A 175, 129 共1999兲.
254 285
J. Miragliotta, R. S. Polizotti, P. Rabinowitz, S. D. Cameron, and R. B. A. Peremans et al., Nucl. Instrum. Methods Phys. Res. A 375, 657
Hall, Appl. Phys. A: Mater. Sci. Process. B51, 221 共1990兲. 共1996兲.
255 286
A. Bandara, J. Kubota, K. Onda, A. Wada, S. S. Kano, K. Domen, and C. Z. Chen, R. Ward, Y. Tian, S. Baldelli, A. Opdahl, Y. R. Shen, and G. A.
Hirose, Surf. Sci. 428, 331 共1999兲. Somorjai, J. Am. Chem. Soc. 122, 10615 共2000兲.

J. Vac. Sci. Technol. A, Vol. 19, No. 6, NovÕDec 2001

Redistribution subject to AVS license or copyright; see http://scitation.aip.org/termsconditions. Download to IP: 128.138.73.68 On: Sun, 21 Dec 2014 09:13:30

You might also like