Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

PC66CH09-Wang ARI 28 February 2015 12:19

Quantitative Sum-Frequency
ANNUAL
REVIEWS Further Generation Vibrational
Click here for quick links to
Annual Reviews content online,
including:
Spectroscopy of Molecular
• Other articles in this volume
• Top cited articles Surfaces and Interfaces:
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

• Top downloaded articles


• Our comprehensive search
Lineshape, Polarization,
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

and Orientation
Hong-Fei Wang,1 Luis Velarde,2 Wei Gan,3
and Li Fu1
1
William R. Wiley Environmental Molecular Sciences Laboratory, Pacific Northwest National
Laboratory, Richland, Washington 99352; email: hongfei.wang@pnnl.gov
2
Department of Chemistry, University at Buffalo, The State University of New York, Buffalo,
New York 14260
3
Xinjiang Technical Institute of Physics and Chemistry, Chinese Academy of Sciences, Urumqi,
Xinjiang 830011, China

Annu. Rev. Phys. Chem. 2015. 66:189–216 Keywords


First published online as a Review in Advance on nonlinear susceptibilities, molecular polarizability, Euler transformation,
December 8, 2014
interference, Fresnel factor, local field factor
The Annual Review of Physical Chemistry is online at
physchem.annualreviews.org Abstract
This article’s doi: Sum-frequency generation vibrational spectroscopy (SFG-VS) can provide
10.1146/annurev-physchem-040214-121322
detailed information and understanding of the molecular composition, in-
Copyright  c 2015 by Annual Reviews. teractions, and orientational and conformational structure of surfaces and
All rights reserved
interfaces through quantitative measurement and analysis. In this review, we
present the current status of and discuss important recent developments in
the measurement of intrinsic SFG spectral lineshapes and formulations for
polarization measurements and orientational analysis of SFG-VS spectra.
The focus of this review is to present a coherent description of SFG-VS and
discuss the main concepts and issues that can help advance this technique as a
quantitative analytical research tool for revealing the chemistry and physics
of complex molecular surfaces and interfaces.

189
PC66CH09-Wang ARI 28 February 2015 12:19

1. INTRODUCTION
In this review, we discuss recent developments in the quantitative measurement and analysis of
SFG-VS: surface sum-frequency generation vibrational spectroscopy (SFG-VS), mainly by connecting the
sum-frequency following three concepts: the spectral lineshape, polarization of light, and molecular orientation.
generation vibrational The overall goal is to provide physical insights through an operable framework for quantitative
spectroscopy
SFG-VS studies and applications. As SFG-VS is a large and rapidly growing area of experimental
Spectral lineshape: and theoretical research, it cannot be covered completely in one review. We have chosen to limit
the form of a feature
our focus to topics not explicitly covered by many prior reviews on the subject, especially those
observed in
spectroscopy usually published in this journal in the past two decades (1–9).
described by SFG-VS is a coherent, second-order nonlinear optical process in which a visible photon (ω1 )
parameters such as the and an infrared (IR) photon (ω2 ) combine into one new photon at the sum frequency of the two
peak position, incident photons (ω = ω1 + ω2 ) through simultaneous interaction with a nonlinear optical medium
maximum height, and
(10). Symmetry requirements for second-order nonlinear optical processes dictate that the SFG-
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

half width
VS signal from the surface of an isotropic medium or the interface between two isotropic media
Polarization:
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

is dominated by surface or interfacial dipolar contributions (11). This gives SFG-VS its intrinsic
the direction of the
electric field of a surface/interface selectivity and makes it a unique spectroscopic tool for molecular surface and
propagating light wave interface studies. By tuning the visible or IR frequencies, investigators can use SFG-VS to probe
Isotropic: uniform in the electronic and vibrational spectra of the molecular surface and interface, respectively (12), and
all orientations or to provide chemically specific information.
directions Historically, theoretical and experimental work on surface SFG-VS was a natural extension
Dipole: a separation of developments in surface second harmonic generation (SHG) (13). Soon after the initial ex-
of positive and perimental realization and theoretical understanding of surface SHG (14–16) and SFG-VS (17,
negative charges 18) in the 1980s, these techniques grew into powerful and versatile in situ spectroscopic tools for
molecular surface and interface studies (1, 12), with applications to almost all kinds of surfaces
and interfaces accessible by light (i.e., gas/liquid, gas/solid, liquid/solid, liquid/liquid, and even
solid/solid interfaces) (19). Among these applications, liquid interfaces (20–23), such as diverse
aqueous surfaces (2, 24–27) and colloidal, biological, and nanoparticle surfaces buried in the so-
lution phase (7, 28, 29), are particularly important and exciting, as they had not been explicitly
studied by a spectroscopic technique, given that the surface signals are largely overwhelmed by
bulk contributions (21). In addition, SFG-VS applications to catalytic (30, 31), polymeric (3, 32),
and biomacromolecular (33–35) surfaces and interfaces have also proven the technique’s versatility
and importance in material, chemical, and biological sciences.
The success of SFG-VS in the past 30 years (36) has resulted from advances in quantitative mea-
surements and our understanding of the SFG-VS spectral lineshape and polarization dependences.
Through them, we obtain detailed orientational and conformational structural information, as well
as information about the dynamic interactions of molecular surfaces and interfaces that other tech-
niques have difficulty with or cannot obtain. Recent developments in SFG instrumentation and
theoretical formulation have offered great promises in transforming SFG-VS from a qualitative
research tool into a set of quantitative analytical and research tools (34, 36–42). To review these
developments with the aim of providing researchers in different fields a well-defined technique
and set of operational tools, we first briefly discuss the issue of the surface dipole contribution and
surface selectivity of SFG-VS; then we present formulations on the quantitative spectral lineshape,
polarization-dependent measurements, and orientational analysis, with references to examples in
the recent literature.

190 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

2. SURFACE DIPOLE CONTRIBUTION AND SURFACE


SELECTIVITY OF SFG-VS
The quantitative analysis of SFG-VS is formulated based on the assumption that the origin of the Quadrupole:
SFG-VS signals can be attributed to surface dipole contributions, which has been confirmed by a a sequence of
large body of experimental and theoretical work (11, 12, 21, 23, 43). alternating positive
and negative charges,
as arranged on the
2.1. Surface Dipole Contribution to SHG and SFG-VS corners of a square

A long-standing question of debate has been how much of the total surface SHG or SFG-VS signal Centrosymmetry:
a point group that
originates purely from the surface dipole contributions and how much originates from the bulk
contains an inversion
multipolar contributions near the surface or interface region of a centrosymmetric or isotropic center as one of its
medium (11, 44–49). The most optimistic picture based on quantum electrodynamics theory symmetry elements
suggested that there is no coherent dipole and multipolar contribution from the isotropic liquid
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

to SHG (and SFG-VS) (44). However, because of the field discontinuity or field gradient across
the surface region, it was also recognized that there is a mechanism for the quadrupolar terms to
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

contribute to the surface SHG signal, even for an isotropic liquid (45). The main difficulty is that
the magnitude of the quadrupolar contribution cannot be quantitatively evaluated, and there is no
simple way to experimentally separate it from the pure surface dipole contributions (11, 50).
Early experimental evidence for a significant quadrupolar contribution was based on the
substantial temperature dependence and breaking of the so-called Kleinman symmetry of the
SHG signal from the air/water interface measured in the reflection geometry (51). From this
study, the ∼35–60% drop in the total SHG signal measured in different polarizations as the
temperature increased from 9◦ C to 80◦ C was attributed to the disappearance of the dipolar
contribution, and the rest was attributed to the bulk quadrupolar contribution. However, later
studies showed that the temperature dependence was an experimental artifact (21, 24, 52).
Moreover, the breaking of the Kleinman symmetry can be successfully treated with purely surface
dipole contributions from the water molecule with C2v symmetry (52). Further examination of
the molecular dispersion relationships and symmetry properties revealed that there is no such
thing as Kleinman symmetry in any nonlinear optical process for a realistic molecule or material
(52–54). In short, so far there has been no experimental evidence for a significant quadrupole
contribution for SHG in the reflective geometry. Claims for such attributions were based either
on flawed experimental data (51, 55, 56) or on an oversimplification of the molecular symmetry
properties using the nonexistent Kleinman symmetry argument (52, 57).
Shen and coworkers (11, 47–49) showed that the SFG-VS signal detected in the reflection
geometry does not have a detectable bulk multipolar contribution, whereas the signal detected in
the transmission geometry might have significant bulk multipolar contributions. The argument
was based on the rationale that the coherence length of the SFG-VS (or SHG) process is much
longer in the transmission direction than in the reflection direction, and this coherence length
determines the number of bulk molecules near the surface or interface that can make multipolar
contributions to the surface SHG or SFG-VS signal. Nevertheless, one has to realize that the
coherence-length argument sets an upper limit for the bulk quadrupolar contributions to the
surface SFG-VS signal. Future detailed experimental and theoretical studies may likely further
demonstrate that the actual bulk quadrupolar contribution is also much smaller than this upper
limit for the transmission geometry.
A significant SFG-VS signal of the C–H stretching modes from the air/liquid benzene inter-
face was reported in the reflective geometry (58); therefore, it was not likely from bulk quadrupo-
lar contributions (11). The benzene molecule at equilibrium has D6h symmetry, which is with
centrosymmetry. Thus, the D6h benzene molecule at an interface should not give a detectable

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 191


PC66CH09-Wang ARI 28 February 2015 12:19

IN SITU STUDY OF BIOLOGICAL AND MEMBRANE PROCESSES

Proteins, DNA, lipids, and sugars are the building blocks of life, and they are all chiral molecules or macromolecules.
A biological membrane is where important biological structures exist and functions occur. Chiral SFG-VS is a unique
tool to probe chiral molecular structures at interfaces. As an in situ surface-specific spectroscopic method (34), SFG-
VS can be used to follow the structural changes of chiral and achiral molecules in biological membrane processes.
A detailed understanding of the SFG-VS responses from the intrinsic and orientational chirality of molecules and
molecular assembly at surfaces has allowed the technique to become a promising tool in biological membrane studies
(5, 34, 115). Biomacromolecules are complex, and their interactions are complicated, and with the development of
HR-BB-SFG-VS, it is expected that research in this direction will expand and grow rapidly in the near future.
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

SFG signal. One recent computational study suggested that the observed SFG signal was from
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

both the breaking of the centrosymmetry of the interfacial benzene molecules and the quadrupo-
lar contributions of the benzene bulk liquid phase (59). Because the computation of the bulk
quadrupolar contribution to the surface signal has been known to be difficult, it is likely that the
quadrupolar contribution was overestimated.
In practice, one should always be cautious when drawing conclusions relying on the bulk
quadrupolar contribution when the SFG-VS or SHG data cannot be easily explained by models
that may be oversimplified or nonrealistic. There are plenty of examples in the literature showing
that when the data are not definitive and the math is not rigorous, invoking complicated higher-
order terms is seldom appropriate.

2.2. Surface Selectivity of Chiral SFG-VS


Chiral SFG-VS probes chiral-specific responses of the molecular interface (34, 35) (see the sidebar
In Situ Study of Biological and Membrane Processes). However, it is known that a chiral liquid
also generates an SFG-VS signal (60, 61). Therefore, the surface selectivity of chiral SFG-VS has
been questioned (62).
In the early 2000s, systematic experimental and theoretical studies on chiral SFG-VS from
chiral liquids revealed that the chiral SFG-VS effect is much smaller than initially determined in
the 1960s (63–67). The surface chiral SFG signal was also considered too weak to be observed
(62), unless the electronic resonance condition is satisfied (68). Thus, some initial reports claiming
observation of the surface chiral SFG signal were questioned and debated (62, 69).
However, the weak bulk chiral SFG signal is not a result of each chiral susceptibility tensor
Chiral: refers to a element being small. It is the result of the almost-perfect cancellation between the different chiral
molecule, or relating tensors in the isotropic ensemble average of the isotropic chiral liquid (5, 34). Nevertheless, the
to a molecule, that is molecular surface is always partially ordered, and chiral SFG from chiral liquid surfaces does not
not superimposable on
have such cancellation as in the isotropic chiral liquid (5, 34, 62, 70). Therefore, surface chiral
its mirror image
SFG is not only allowed, but also comparable to the surface achiral SFG signal at the chiral liquid
HR-BB-SFG-VS:
surface (5, 34). Therefore, chiral SFG is also surface selective (5, 34). One recent experiment with
high-resolution
broadband high-resolution broadband sum-frequency generation vibrational spectroscopy (HR-BB-SFG-
sum-frequency VS) on surface achiral and chiral SFG-VS of the R-(+)-limonene and S-(−)-limonene liquid
generation vibrational surfaces provided a clear demonstration of the surface selectivity of chiral SFG-VS, along with a
spectroscopy detailed analysis of the intrinsic chirality and orientational (prochiral) chirality contributions in
the observed SFG signal (71).

192 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

3. QUANTITATIVE MEASUREMENT AND ANALYSIS


OF SFG LINESHAPES
A spectral lineshape is the form of the spectral features observed in a spectroscopic measurement Fourier transform
that contains all the information the spectrum can give. The focus of spectroscopic studies is to convolution
obtain and understand an accurate spectral lineshape and to capture its change in the frequency theorem: under
suitable conditions,
or time domain as the system is transformed. SFG-VS is no exception.
the Fourier transform
of a convolution is the
pointwise product of
3.1. Unified Treatment of Frequency- and Time-Domain SFG-VS Fourier transforms
SFG-VS is a second-order nonlinear molecular response to the simultaneous interacting visible
and IR fields. The response function can be defined either in the time domain or in the frequency
domain, and they are equivalent to each other and are connected through a time-frequency Fourier
transformation (72–75).
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

In the time domain, the SFG emitting field is proportional to the SFG polarization P (2) (t; τ ),
 ∞  ∞
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

P (2) (t; τ ) = dt2 χ (2) (t2 , t1 ; τ )Evis (t − t2 )EIR (t − t2 − t1 )dt1 . (1)


0 0

Here χ (t) is the time-dependent SFG response function, and Evis (t) and EIR (t) are the local
(2)

electric fields of the visible and IR laser lights. P (2) (t) depends parametrically on τ , the time delay
between the arrival of the IR and VIS pulses. t1 and t2 are the integral variables of the IR and
visible light-matter interaction time, respectively.
In the SFG-VS process, the IR excitation is followed by the slower vibrational coherence of the
molecular vibrational states, and the visible response is essentially instantaneous (73, 74). One has
 
P (2) (t; τ ) = Evis (t; τ ) χ (2) (t) ⊗ EIR (t) , (2)
where ⊗ is the convolution operator. The frequency-domain SFG polarization P̃ (2) (ωSF ; τ ) is
simply the Fourier transformation of the time-domain polarization P (2) (t; τ ); that is,
 ∞
1
P̃ (2) (ωSF ; τ ) = e iωt P (2) (t; τ )dt
2π −∞
 ∞
1   (3)
= e iωt Evis (t; τ ) χ (2) (t) ⊗ EIR (t) dt
2π −∞
 
= Ẽvis (ωvis ; τ ) ⊗ χ (2) (ωSF ) ẼIR (ωIR ) .
Here ωSF = ωvis + ωIR is the sum frequency of the visible and IR frequencies, and Ẽvis (ωvis ; τ )
and ẼIR (ωIR ) are the corresponding visible and IR optical fields in the frequency-domain repre-
sentation, respectively. The tilde symbol indicates that the field is a function of the frequency,
and τ represents the time delay between the visible and IR fields. With the Fourier transform
convolution theorem, the convolution in the time domain becomes a simple product in the
frequency domain, and vice versa.

3.2. Intrinsic Lineshape Measurements in Frequency-Domain SFG-VS


SFG intensities in the time and frequency domains are
 ∞
 (2)   
ISFG (τ ) ∝  P (t; τ )2 dt and ISFG (ωSF ; τ ) ∝  P̃ (2) (ωSF ; τ )2 . (4)
−∞

SFG-VS measurements can be conducted either by measuring the intensity (homodyned) or


the field (phase resolved, phase sensitive, or heterodyned) in the time domain, also called SFG

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 193


PC66CH09-Wang ARI 28 February 2015 12:19

(2) ∞
free-induction decay [i.e., ISFG (τ ) (76–78) or ESFG (τ ) ∝ P (2) (τ ) = −∞ P (2) (t; τ )dt (74), respec-
tively], or by measuring the intensity or the field in the frequency domain [i.e., ISFG (ωSF ; τ ) or
(2)
Spectral resolution: ẼSFG ∝ P̃ (2) (ωSF ; τ ) (8, 9), respectively].
a measure of the ability The best way to accurately measure the SFG-VS lineshape function is through HR-BB-SFG-
to resolve features in VS in the frequency domain, that is, through the intensity ISFG (ωSF ; τ ) or through the phase-
the electromagnetic (2)
resolved field ẼSFG ∝ P̃ (2) (ωSF ; τ ) (73, 79, 80). In HR-BB-SFG-VS, the nearly transform-limited
spectrum
visible laser field is very narrow in frequency; therefore, it can be represented as Ẽvis (ωvis ; τ ) =
Evis (ωvis ; τ )δ(ω). Then the convolution operation between the visible field and the rest of the
expression becomes a simple product operation:

Ẽvis (ωvis ; τ ) ⊗ [χ (2) (ωSF ) ẼIR (ωIR )] = χ (2) (ωSF )Evis (ωvis ) ẼIR (ωIR ) (5a)

or
     
 Ẽvis (ωvis ; τ ) ⊗ [χ (2) (ωSF ) ẼIR (ωIR )]2 = χ (2) (ωSF )2  Evis (ωvis ; τ ) ẼIR (ωIR )2
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

 2 (5b)
= χ (2) (ωSF ) I (ωvis ) I˜ (ωIR ).
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

Here the parameter τ is dropped because the visible pulse is so broad in time that the overall
functional form no longer depends on the delay between the visible and IR pulses. A good con-
firmation has been provided by an HR-BB-SFG-VS measurement with an ∼90-ps-long visible
pulse and an ∼100-fs IR pulse in which a delay of 10 ps or more caused no detectable change in
the measured SFG spectral lineshape (73, 80).
χ (2) (ωSF ) [or |χ (2) (ωSF )|2 ] can be obtained directly through normalization over the
Evis (ωvis ) ẼIR (ωIR ) [or I (ωvis ) I˜ (ωIR )] term in Equations 5a,b, as it can be easily measured from
the SFG signal from a reference surface or bulk material with a flat spectral response (e.g., a z-cut
quartz or a gold thin film frequently used in SFG-VS measurements). Thus, the lineshape function
χ (2) (ωSF ) [or its absolute square |χ (2) (ωSF )|2 ] can be accurately measured.
In the SFG literature (38, 41, 43, 81), the SFG intensity has always been written in the form
of |χ (2) (ωSF )|2 I (ωvis )I (ωIR ), and the |χ (2) (ωSF )|2 spectra are proportional to ISFG (ω)/I (ωvis )I (ωIR ).
Implicitly, in those derivations, both the visible and IR lasers are assumed monochromatic. Now,
one can see from Equation 5 that only the visible pulse has to be sufficiently monochromatic, so it
can be well approximated as a δ function in the frequency domain, and there is no requirement on
the monochromaticity for the IR field. This fact enables the use of BB-SFG-VS with a broadband
IR pulse and monochromic visible pulses (37).
In time-domain SFG–free-induction decay measurements, as shown in Equation 2, the con-
volution is on the EIR (t) term. Thus, the IR pulse has to be extremely short to separate the
response function χ (2) (t) that is convoluting with the IR field term, EIR (t). Furthermore, because
∞ ∞ 2
the measurement also involves time averaging −∞ P (2) (t; τ )dt or −∞ t|P (2) (t; τ )| dt, there is not
a simple way to obtain the χ (2) (t) lineshape function through direct normalization of the laser
profiles unless both the visible and IR pulses are δ functions in the time domain. Because neither
the visible nor IR pulses can be much shorter than the molecular vibration cycle, in the time
domain, deconvolution procedures with well-controlled and well-characterized laser profiles are
always required to retrieve the χ (2) (t) lineshape (74, 77, 82). However, the mathematical equations
involved here clearly show that complete retrieval of the SFG lineshape function χ (2) (t) is difficult
or even impossible, especially when there are multiple overlapping peaks.

3.3. How Much Spectral Resolution Is Needed?


It is necessary to evaluate the effects of spectral resolution on the broadening of SFG spectra
and lineshape changes. A simple formula with an accuracy of 0.02% can be used to evaluate the

194 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

Table 1 Broadening of Lorentzian spectral lineshapes, νL , by Gaussian visible probes, νG
νV − ν L (cm−1 )
νG (cm−1 ) νL = 5 cm−1 νL = 10 cm−1 νL = 20 cm−1 νL = 50 cm−1
0.6 0.08 0.04 0.02 0.01
1 0.21 0.11 0.05 0.02
2 0.74 0.41 0.21 0.09
4 2.30 1.48 0.82 0.34
6 4.11 2.94 1.77 0.76
8 6.00 4.60 2.97 1.34
10 7.94 6.38 4.35 2.06
15 12.85 11.05 8.34 4.42

νV − νL values with less than a 10% change from νL are in blue.
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

resulting Voigt line width [full width at half maximum (FWHM) = νV ] from a peak with a
typical Lorentzian lineshape (FWHM = νL = 2 q ) measured by a laser pulse with a typical
Gaussian spectral profile (FWHM = νG ) (73, 83),

νV = 0.5346νL + 0.2166(νL )2 + (νG )2 . (6)
Table 1 lists typical values of νV − νL calculated with different spectral widths, νL , and
spectral resolutions, νG . The typical resolution of scanning SFG-VS (6–10 cm−1 ) and conven-
tional BB-SFG-VS (10–15 cm−1 ) would cause significant broadening of peaks with the relatively
narrower spectral width. For example, when νL = 10 cm−1 (i.e., q = 5 cm−1 ), the 6-cm−1
spectral resolution makes νV − νL = 2.94 cm−1 , a nearly 30% broadening. In addition to
the spectral broadening, there is a change in the spectral lineshape from Lorentzian to Voigt,
assuming the laser profile is Gaussian. For a known isolated spectral peak, the broadening and
lineshape change probably can be corrected with known visible and IR laser profiles. However,
the spectral broadening and lineshape change can introduce errors and uncertainties in analyzing
or fitting spectra with closely overlapping peaks (27, 84). In some SFG-VS experiments that use
Voigt lineshape:
asymmetric visible laser profiles (85–87), artifacts such as lineshape distortion and phase flipping in
a convolution of the
the SFG-VS spectra can be observed (74, 88). According to Table 1, the instrumental resolution Lorentzian and
should usually be at least 2 cm−1 , unless all the SFG-VS spectral features are known to have widths Gaussian lineshapes
larger than 10 cm−1 . Lorentzian
lineshape:
a continuous
3.4. Parameters from Intrinsic SFG-VS Lineshapes probability
distribution associated
The intrinsic time-domain response function χ (2) (t), which contains all the molecular vibrational
with homogeneous
coherence information, including both the homogeneous (T 2 ) and inhomogeneous (ω) broad- broadening
ening parameters, is a solution to the optical Bloch equations and is well approximated as (73, 74,
Gaussian lineshape:
89, 90) a continuous
 2 2
χ (2) (t) = |ANR | e iψNR δ(t) − iθ(t) Aq e −iωq t e −t/T 2q e −ωq t /2 . (7) probability
q distribution usually
associated with
(Here, ωq is the resonance frequency of the q-th mode in angular frequency units, T 2q is the
inhomogeneous
homogeneous broadening parameter, and ωq is the inhomogeneous dephasing parameter for broadening or laser
the q-th mode. To connect to the commonly used spectroscopic values, one has ωq = 2π c ν̃q width broadening
and ωq = 2π c δ ν̃q , where ν̃q is the q-th vibrational mode frequency in wave numbers, and c is

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 195


PC66CH09-Wang ARI 28 February 2015 12:19

the speed of light in centimeters per second. ANR and ψNR are the amplitude and phase of the
nonresonant term, respectively. δ(t) is the δ function at the zero IR excitation time, and θ(t) is the
Heaviside step function introduced by the causality of the IR interaction.
According to Equation 3, χ (2) (ωSF ) is the Fourier transformation of χ (2) (t) and χ (2) (ωSF ),

ω2
 Aq − IR2
χ (2) (ωIR ) = |ANR | e iψNR + ⊗ e 2ωq
q
ωq − ωIR − i q
 (ω −ω )2
(8)
 ∞ Aq − IR 2
= |ANR | e iψNR
+ e 2ωq d ω .
q 0 ωq − ω − i q

Here, the variable ωSF is replaced by ωIR because as the visible frequency ωvis in SFG-VS is usually
fixed, the change in ωSF is the same as the change in ωIR . Thus, Equations 7 and 8 provide a
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

unified description of the SFG response in the time and frequency domains, respectively. Because
the molecular parameters in Equations 7 and 8 are identical, measurement of either χ (2) (ωSF ) or
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

χ (2) (t) provides the same information for the molecular system. As discussed above, χ (2) (ωSF ) or
|χ (2) (ωSF )|2 can be accurately obtained from frequency-domain SFG-VS experiments as long as the
visible pulse is spectrally narrow; thus, we focus the discussion on frequency-domain measurements
and analysis.

3.5. Sub-Wave-Number HR-BB-SFG-VS


In 2011, a sub-wave-number HR-BB-SFG-VS spectrometer using a synchronized 90-ps visi-
ble laser pulse and sub-100-fs IR pulse achieved 0.6-cm−1 resolution (79). Subsequently, sev-
eral benchmark systems have been studied to illustrate this technique’s ability to obtain nearly
intrinsic SFG-VS spectral lineshapes. For example, overlapping peaks have been resolved as
close as 2.8 cm−1 from a seemingly single spectral peak with 8.6-cm−1 FWHM (79), and ac-
curate Voigt lineshape parameters of –CN vibrations in a monolayer with both homogeneous
and inhomogeneous broadening have been obtained (80). Twelve overlapping peaks of the C–
H stretching vibration spectra of a cholesterol monolayer at the air/water interface (Figure 1)
have been resolved, and time-dependent SFG-VS spectra with lower resolution under time delay
conditions have been successfully predicted, with the parameters obtained from high-resolution
spectra (76). Additionally, achiral and chiral spectra of a chiral liquid surface have been de-
termined, aiding our understanding of the chirality and prochirality of a chiral interface (71).
These studies not only proved the validity of the treatment described above, but also demon-
strated how this new tool can take SFG-VS quantitative measurements and analysis to a dif-
ferent level for applications to study complex structures and interactions at the surface and
interfaces.

3.6. Spectral Lineshapes in Intensity and Phase-Resolved SFG-VS


Recently, phase-sensitive or phase-resolved SFG-VS has gained much popularity in SFG-VS
studies (8, 9, 91) through many applications, particularly in measuring O–H stretching vibra-
tional spectra at various aqueous interfaces (92–98). Phase-resolved SFG-VS measurements are
important because the relative phase of different spectral features can be directly related to the
relative or even absolute orientations of the corresponding molecular groups. Therefore, there
have been many discussions on the advantages and limitations of phase-resolved SFG-VS and
intensity SFG-VS (8, 9, 91, 99, 100).

196 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

0.25
a b
SFG resolution: 0.20 90-ps Vis,
0.20 HR-BB-SFG (0.6 cm –1) IR-Vis delay (τ):

Normalized intensity
SFG intensity (a.u.)

BB-SFG (15 cm –1) 0.15 Zero


0.15 ps scanning (6 cm –1) 10 ps

0.10
0.10

0.05 0.05

0.00 0.00
2,750 2,800 2,850 2,900 2,950 3,000 3,050 2,750 2,800 2,850 2,900 2,950 3,000 3,050
IR frequency (cm –1) IR frequency (cm –1)
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

Figure 1
Intrinsic spectral lineshape measured with high-resolution broadband sum-frequency generation (HR-BB-SFG) vibrational
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

spectroscopy on a cholesterol Langmuir monolayer (35 Å2 per molecule surface density) on an air/water interface. (a) Comparison of
SFG with spectral resolutions of 0.6 cm−1 from HR-BB-SFG, 6 cm−1 from picosecond scanning SFG (ps scanning), and 15 cm−1 from
conventional broadband SFG (BB-SFG). The 0.6-cm−1 resolution spectrum allows twelve peaks to be uniquely and accurately
resolved. (b) Comparison of a 10-ps time delay and a zero delay between the visible (Vis) and IR pulse in HR-BB-SFG using a 90-ps
visible pulse, showing essentially no spectral lineshape and intensity change, as in the discussion associated with Equation 5. Figure
reproduced from Reference 73 with permission from the PCCP Owner Societies.

Ideally, the relative phase of overlapping peaks should be revealed in the detailed lineshape
in their intensity spectra through spectral interference. However, an accurate lineshape of the
intensity spectrum is usually unavailable, and was especially difficult to obtain prior to the devel-
opment of HR-BB-SFG-VS. The measured spectra might give uncertain relative phases to the
overlapping spectral features with spectral fitting (9, 27, 84, 101). Lineshape retrieving techniques,
such as the maximum entropy method, have been employed to help resolve the relative phases
between overlapping spectral features in the intensity spectrum (84, 101, 102).
Phase-resolved SFG-VS measurements have also been used to address this problem (9). How-
ever, a phase-resolved SFG-VS experiment is usually more complicated to implement than its
intensity SFG-VS counterpart (9). There are many practical issues involved in obtaining good
lineshapes for phase-resolved spectra (99, 100). Therefore, the interpretation of phase-resolved
SFG-VS spectra has been mostly qualitative instead of quantitative. Recent developments in HR-
BB-SFG-VS suggest that intensity SFG measurements with high spectral resolution are likely to
have an advantage given their simplicity and ability to obtain nearly intrinsic spectral lineshapes
(80). However, for spectra with complicated lineshapes, such as O–H spectra from the aqueous
interface, phase-resolved measurements have been of great value (9).
One clear advantage of phase-resolved SFG-VS measurements is in pump-probe measurements
using the heterodyne technique (74, 103, 104). In a pump-probe experiment, the spectral resolution
is sacrificed to some extent for time resolution. Therefore, spectral lineshapes at different pump-
probe delay times are complicated because of interference, and it is difficult to analyze time-
dependent spectral changes in the intensity data. In contrast, phase-resolved SFG-VS data using
the heterodyne technique are not subject to spectral interference, and the measured spectral
amplitude is linearly proportional to the surface density Ns . This can be of great advantage in
following the dynamics of population changes. However, because the spectral resolution and
lineshape accuracy are usually sacrificed, caution is needed in analyzing the spectra in detail.

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 197


PC66CH09-Wang ARI 28 February 2015 12:19

4. POLARIZATION MEASUREMENT AND ORIENTATION


ANALYSIS IN SFG-VS
Around 2000, several papers and two student theses from Ron Shen’s group laid out a detailed
formulation for quantitative analysis with SFG-VS and SHG signals measured in different po-
larization combinations (38, 39, 105, 106). These works summarized and further developed the
formulations toward SFG-VS quantitative analysis reported in the previous decade (107–112).
These formulations were further advanced and tested with quantitative polarization-dependent
measurements and analysis in connection with detailed molecular symmetry analysis (34, 40, 41,
113). The main formulations are summarized below.

4.1. Overall Expressions of SFG-VS


In the frequency domain, the SFG-VS intensity, measured with a spectrally narrow visible pulse,
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

can be written as (38, 41)


 
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

8π 3 ω2 sec2 β  (2) 2
I (ω) = 3 χeff  I (ω1 )I (ω2 ), (9)
c nI (ω)nI (ω1 )nI (ω2 )
with
(2)
χeff = [ê(ω) · L(ω)] · χ(2) : [L(ω1 ) · ê(ω1 )] [L(ω2 ) · ê(ω2 )] . (10)
Here ω, ω1 , and ω2 are the frequencies of the SFG signal, visible, and IR laser beams, respectively.
nm (ωi ) is the refractive index of the distinct bulk or interface medium m (m = I, II, and  for the
incident bulk medium, the second bulk medium, and the interface, respectively) at frequency ωi
(i = 0, 1, and 2); βi is the incident or reflection angle from the interface normal to the beam with
frequency ωi ; I (ωi ) is the intensity of the SFG signal or the input laser beams; and ê(ωi ) is the unit
electric field vector for the light beam at frequency ωi at the interface. The SFG signal direction
angle β (Figure 2) should satisfy the relationship ω sin β = ω1 sin β1 + ω2 sin β2 (114). χ(2) is the
frequency-domain SFG response function, also called the SFG-VS susceptibility, as discussed in

Ep
ê(ω)
Ω2 z
Es ê(ω2) Ep Ω ω
Ep
Ω1 ω2 β1 Es
Es ê(ω1) β2
β Interface
ω1
Phase I

Phase II

z
γ2
y γ1
γ ω1

x ω
ω2

Figure 2
Illustration of the experimental configuration of sum-frequency generation vibrational spectroscopy in the
common copropagating reflection geometry.

198 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

Section 3.1. It is a third-rank tensor with 3 × 3 × 3 = 27 tensor elements, which are usually
represented as χi j k , with (i, j, k) as the indexes, each representing one of the three axes of the
laboratory coordinate system (x, y, and z).
Equations 9 and 10 give the full expression of the SFG-VS intensity that can be derived by
inserting the SFG polarization P̃ (2) (ωSF ) as the SFG field radiation source term into the Maxwell
equation with proper boundary conditions (43, 81). They can also be derived by the microscopic
theory of SFG-VS based on the Ewald-Oseen summation of the microscopic radiation fields of a
lattice of discrete surface nonlinear dipoles (42).
Most SFG-VS experiments are performed with a copropagation reflective geometry, as de-
scribed in Figure 2, in which the incident angles, polarization angles of the SFG, and the visible
and IR beams are defined in the laboratory coordinate system (x, y, and z). It had been shown
previously that the copropagation geometry has certain advantages in polarization analysis and
orientation angle calculations (41). As discussed in Section 2, in the reflective geometry, the de-
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

tectable SFG-VS signal comes exclusively from the surface dipole contribution. Therefore, the
formulation below concerns only the surface dipole contributions to χ(2) .
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

With the defined laboratory coordinate system, the 27 tensor elements of χi j k for a rotational
isotropic surface along the interface normal direction have only 13 nonzero elements (i.e., the seven
achiral terms χxxz = χ y y z , χxzx = χ y zy , χzxx = χzy y , and χzzz and the six chiral terms χxy z = −χ y xz ,
χzxy = −χzy x , and χxzy = −χ y zx ) (40). The chiral terms always have i = j = k, and the achiral
terms always have at least two of the three indexes the same. In the past, the achiral terms and
chiral terms were usually discussed separately. However, as it is now known that achiral molecules
can form orientational chiral structures at an interface (5, 40, 113), a unified formulation with
both the achiral and chiral terms is necessary.
With knowledge of the nonzero χi j k tensors, and the definitions of the incident angle (βi ) and
polarizations angles ( i ) as in Figure 2, the unit electric field vector ê(ωi ) in Equation 10 can be
(2) (2) (2)
projected onto the (x, y, z) axis, and the detailed expressions of χeff = χeff,achiral + χeff,chiral are (34,
115, 116)
(2)
χeff,achiral = sin sin 1 cos 2 L y y (ω)L y y (ω1 )L zz (ω2 ) sin β2 · χ y y z
+ sin cos 1 sin 2 L y y (ω)L zz (ω1 )L y y (ω2 ) sin β1 · χ y zy
+ cos sin 1 sin 2 L zz (ω)L y y (ω1 )L y y (ω2 ) sin β · χzy y
− cos cos 1 cos 2 L xx (ω)L xx (ω1 )L zz (ω2 ) cos β cos β1 sin β2 · χxxz (11a)
− cos cos 1 cos 2 L xx (ω)L zz (ω1 )L xx (ω2 ) cos β sin β1 cos β2 · χxzx
+ cos cos 1 cos 2 L zz (ω)L xx (ω1 )L xx (ω2 ) sin β cos β1 cos β2 · χzxx
+ cos cos 1 cos 2 L zz (ω)L zz (ω1 )L zz (ω2 ) sin β sin β1 sin β2 · χzzz ,

(2)
χeff,chiral = sin cos 1 cos 2 L y y (ω)L zz (ω1 )L xx (ω2 ) sin β1 cos β2 · χ y zx
+ sin cos 1 cos 2 L y y (ω)L xx (ω1 )L zz (ω2 ) cos β1 sin β2 · χ y xz
+ cos cos 1 sin 2 L zz (ω)L xx (ω1 )L y y (ω2 ) sin β cos β1 · χzxy
(11b)
− cos cos 1 sin 2 L xx (ω)L zz (ω1 )L y y (ω2 ) cos β sin β1 · χxzy
+ cos sin 1 cos 2 L zz (ω)L y y (ω1 )L xx (ω2 ) sin β cos β2 · χzy x
− cos sin 1 cos 2 L xx (ω)L y y (ω1 )L zz (ω2 ) cos β sin β2 · χxy z .
In these expressions, the negative sign comes from the fact that the x projection of the SFG field
is opposite to the x projections of the visible and the IR fields in the copropagation geometry.
Thus, one can easily find the correct expressions by switching the proper signs in Equations 11a,b
accordingly for the counterpropagating geometry or other geometries.

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 199


PC66CH09-Wang ARI 28 February 2015 12:19

4.2. Fresnel Factors


The SFG Fresnel factors in Equation 10 should not be confused with the Fresnel factors for linear
Fresnel factor: reflections, which are (39, 42, 106)
a factor that describes
how light is reflected 2εI (ωi )kIIz (ωi ) 2nI (ωi ) cos γi
L xx (ωi ) = = , (12a)
from each smooth εII (ωi )kIz (ωi ) + εI (ωi )kIIz (ωi ) nI (ωi ) cos γi + nII (ωi ) cos βi
surface as a function of
the incidence angle 2kIz (ωi ) 2nI (ωi ) cos βi
and wavelength L y y (ωi ) = = , (12b)
kIz (ωi ) + kIIz (ωi ) nI (ωi ) cos βi + nII (ωi ) cos γi
Anisotropic:
not uniform in all 2εI (ωi )εII (ωi )kIz (ωi ) 2n2I (ωi )nII (ωi ) cos βi
orientations or L zz (ωi ) = = . (12c)
εII (ωi )kIz (ωi ) + εI (ωi )kIIz (ωi ) nI (ωi ) cos γi + nII (ωi ) cos βi
directions
Local field factor: Here, εI (ωi ) = n2I (ωi ) and εII (ωi ) = n2II (ωi ) are the dielectric constant of the upper and lower
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

factor that determines phases, respectively; γi is the refraction angle that satisfies nI (ωi ) sin βi = nII (ωi ) sin γi ; and kIz (ωi )
the field experienced and kIIz (ωi ) are the projection of the wave vector of the i-th field onto the z direction in phase I
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

by an atom or and II, respectively.


molecule of interest
from the average
Equations 12a–c are derived for a three-layer model, with the interface as the middle layer
macroscopic field with a macroscopic dielectric constant equal to 1, or the vacuum value. This model can be gen-
erally used for exposed and buried interfaces, such as the solid/liquid interface, and the inverted
gas/solid interface (39, 42, 106). When the εI (ωi ) term on the upper level of L zz (ωi ) is removed,
the expressions in Equations 12a–c are the same as those for the exposed two-layer model (42).
Whether the interface can be treated as a third layer is debatable. However, in most cases, it can be
shown that the two models would essentially end up with the same L zz (ωi ) (42). In the literature,
the L zz (ωi ) term would include an effective dielectric constant ε as

2εII (ωi )kIz (ωi ) εI (ωi ) 2nII (ωi ) cos βi n I (ωi ) 2


L zz (ωi ) = ·  = · . (13)
εII (ωi )kIz (ωi ) + εI (ωi )kIIz (ωi ) ε (ωi ) nI (ωi ) cos γi + nII (ωi ) cos βi n (ωi )
The effective dielectric constant is not the dielectric constant of the interface. It is the ratio of
two anisotropic local field factors ε  (ωi ) = (l  (ωi ))/(l ⊥ (ωi )), with l xx (ωi ) = l y y (ωi ) = l  (ωi ) and
l zz (ωi ) = l ⊥ (ωi ) (38, 39, 42, 117). The definition of the microscopic local field factors in the
molecular layer is not provided in macroscopic surface SHG and SFG theory (39, 42). However,
it can be explicitly defined with molecular polarizabilities and the packing order of the dipoles in the
interface layer using microscopic surface SHG and SFG theory (42). The details of the microscopic
theory are not discussed here. However, it is important to know that the local field factors of the
interface layer are the unique dielectric properties of the interface layer, not a property that can
be averaged from the two bulk layers. The intrinsic anisotropy of the interface layer dictates the
anisotropy of the local field factors in the interface layer; this anisotropy makes the treatment of
interfacial optical and electric properties more difficult but interesting and important. SFG-VS is
sensitive to this anisotropy and therefore can be used to determine it. Finally, different molecular
groups in the same molecule in the interface layer may have different local field factors, and such
details can be studied with SFG-VS (42).

4.3. Interfacial Local Field Factors, Macroscopic Susceptibility,


and Microscopic Polarizability
The macroscopic susceptibility tensor χi j k is the ensemble orientational average of the microscopic
polarizability tensor βi  j  k (also with 27 elements defined in the molecular-frame coordinate sys-
tem, with some being nonzero, depending on the molecular symmetry). Thus, each χi j k is a linear

200 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

combination of the 27 βi  j  k tensor elements through the rotational average and transformation
from the molecular frame to the laboratory frame (111), and the expressions can be significantly
simplified according to the symmetry of the interface, as well as the symmetry of the molecular
Euler angles:
groups under consideration (40). the three angles
One has introduced by

χi j k = N s l ii (ω)l j j (ω1 )l kk (ω2 )
Rii  R j j  Rkk βi  j  k . (14) Leonhard Euler to
describe the
i  j  k
orientation of a rigid
Here N s is the interface molecular number density per unit area; l ii (ωi ), l j j (ωi ), and l kk (ωi ) are the body
microscopic local field factors as l xx (ωi ) = l y y (ωi ) = l  (ωi ) and l zz (ωi ) = l ⊥ (ωi ), discussed above
in associated with Equations 12c and 13. When χi j k is expressed in the form of Equation 14,
the expression of L zz (ωi ) in Equation 12c has to be used. However, as in most polarization
analysis, the individual local field factors are usually not needed. To simplify the analysis, one
usually expresses the χi j k tensor as (38, 39, 42)
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org


χi j k = N s
Rii  R j j  Rkk βi  j  k , (15)
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

i  j  k

and an effective interface dielectric constant ε (ωi ) can be lumped into the definition of L zz (ωi )
as defined in Equation 13 to properly represent the interface anisotropy (38). ε (ωi ) is defined as
(38, 39, 42)
l  (ωi )
ε (ωi ) = . (16)
l ⊥ (ωi )
In this rearrangement, a common factor l  (ω)l  (ω1 )l  (ω2 ) is dropped from the expression of χi j k .
This is not a problem for polarization analysis and comparison of different spectral features of
the same interface. However, as each of the l  (ωi ) factors is a function of the separation and tilt
angle of the dipoles in the interface, the common factor l  (ω)l  (ω1 )l  (ω2 ) needs to be put back into
the analysis when a quantitative comparison of the spectral intensity at different surface densities
and molecular orientations is needed. In short, Equation 14 goes with Equations 12c, 15, and 16,
which go with Equation 13. Equation 14 is the accurate representation, and Equation 15 is in a
simplified form and is often used in SFG-VS analysis (38, 41).

4.4. Euler Transformation and Ensemble Rotational Average


The rotational ensemble average
Rii  R j j  Rkk in Equations 14 and 15 is defined with the matrix
elements of the Euler transformation of the three rotational Euler angles (θ, φ, and ψ) to transform
a property, such as βi  j  k , from the molecular-frame coordinate system to the corresponding
property in the laboratory frame, such as χi j k . Therefore, knowledge of the Euler transformation
is essential to connect the macroscopic measurements to the microscopic molecular properties
(40, 41).
There are 12 different ways (2 × 3! = 12) to perform the Euler transformation (118, pp. 150–
54). The so-called z-x-z convention has been treated as standard in textbooks (118), as it has been
widely used in celestial mechanics, applied mechanics, and frequently in molecular and solid-
state physics, as well as in nonlinear optics (10, 40, 41, 118, 119). Therefore, this convention is
recommended for general use, and we follow it in this review.
The z-y-z convention is also used (34, 120). When switching from the z-x-z convention to the
z-y-z convention, one should use ψ − 90◦ to replace ψ, and 90◦ + φ to replace φ, respectively, in
the z-x-z transformation matrix, and vice versa (41, 118). To avoid confusion caused by mixing up
different conventions or using an erroneous transformation matrix, we recommend that students
and researchers check the validity of the transformation matrix and the definition of the angles

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 201


PC66CH09-Wang ARI 28 February 2015 12:19

before making further derivations or calculations. The standard textbook to consult is Goldstein’s
(118) Classical Mechanics.
Then, the ensemble average is
 π  2π  2π
Rii  R j j  Rkk f (θ, φ, ψ) sin θd θd φd ψ

Rii  R j j  Rkk = 0 0  π 0 2π  2π , (17)
0 0 0
f (θ, φ, ψ) sin θd θd φd ψ
where f (θ, φ, ψ) is the orientational distribution function, and it is usually in a Gaussian form as
(39)

(θ − θ0 )2 (φ − φ0 )2 (ψ − ψ0 )2
f (θ, φ, ψ) ∝ Aexp − − − , (18)
2σθ2 2σφ2 2σψ2
where A is the normalization prefactor; θ0 , φ0 , and ψ0 are the centers of the respective Euler
angles; and σθ , σφ , and σψ are the standard deviations. Usually, f (θ, φ, ψ) is only a function of the
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

tilt angle θ and the twist angle ψ.



There are a few important properties of the ensemble average with Rii  R j j  Rkk (40, 41). First,
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

for a rotationally isotropic surface or interface, the chiral βi  j  k (i  = j  = k ) term can appear
in the achiral χi j k term only in the form of the fully symmetric permutation (34, 40, 115). Their
contributions are usually insignificant except for antiparallel structures, for which the achiral terms
cancel each other in the local structural unit (5, 69). Second, the achiral βi  j  k terms of the molecular
groups with Cs or C2v symmetry can contribute to the chiral χi j k term for a chirally oriented system
but not for a helix structure (34, 71). This is the so-called orientational chirality or prochirality
(2)
(5, 40, 71, 113). Third, for a rotationally isotropic surface or interface, the achiral terms χeff,achiral
have the orientational functional form of N s d eff (
cos θ − c eff ·
cos θ ), an odd function of
cos θ
3
(2)
(41, 121), whereas the chiral terms χeff,chiral have the form of N s d eff,c
sin2 θ , an even function of

cos θ (40, 71). Here, the intensity parameters d eff and d eff,c and the orientational c eff parameter
are functions of the twist angle ψ, as well as the corresponding polarization angles, incident angles
(βi  j  k ), etc. These orientational functional formulations are the basis for quantitative orientational
analysis and the development of polarization selection rules (41, 71, 121).

4.5. SFG-VS Spectral Lineshape and Vibrational Mode-Specific Analysis


According to Equation 8, the SFG-VS susceptibility in its tensorial form can be written as
ω2

χq ,i j k − 22
χi j k = χNR,i j k + ⊗ e 2ωq
q
ωq − ω2 − i q
 (ω −ω )2
(19)
 ∞ χq ,i j k − 2 2
= χNR,i j k + e 2ωq
d ω ,
q 0 ωq − ω  − i
q

and the molecular susceptibility can be written as


ω2
 βq ,i  j  k − 22
β i  j  k =β NR,i  j  k + ⊗ e 2ωq
q
ωq − ω2 − i q
 (ω −ω )2
(20)
 ∞ βq ,i  j  k − 2 2
= βNR,i  j  k + e 2ωq
d ω .
q 0 ωq − ω − i q

These expressions are essentially the same as the lineshape functions used in the SFG-VS
literature for spectral analysis, except that the frequency part is written as ωq − ω2 − i q instead
of ω2 − ωq + i q (38, 41, 73, 80). Because ωq − ω2 − i q = −(ω2 − ωq + i q ), the two forms are

202 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

equivalent, with differences only in the relative sign between the nonresonant term χNR,i j k and the
amplitude terms χq ,i j k . The new ωq −ω2 −i q form keeps the parameters in the frequency-domain
lineshape function (Equation 8) consistent with those in the time-domain lineshape function
(Equation 7) (41). It is suggested that the new forms in Equations 19 and 20 be used for consistency
and to avoid confusion in making coherent dynamics simulations using the parameters from the
fit of SFG spectra.
Inserting Equations 19 and 20 into Equation 15, one has

χq ,i j k = N s
Rii  R j j  Rkk βq ,i  j  k . (21)
i  j  k

Equation 21 gives the vibrational mode–specific expressions of χq ,i j k as a function of the orienta-


tionally averaged tilt (θ) and twist (ψ) angles, and the vibrational mode–specific βq ,i  j  k tensors.
SFG-VS theory demonstrates that the mode-specific polarizability tensor is the product of the IR
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

transition dipole ∂μk /∂ Q q = μk and the Raman polarizability tensor ∂αi  j  /∂ Q q = αi j  as (40,
122–124)
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

1 ∂αi  j  ∂μk
βq ,i  j  k = − . (22)
2ε0 ωq ∂ Q q ∂ Q q
Equation 22 dictates that a vibrational mode observed in SFG-VS spectroscopy has to be both
IR and Raman active. Moreover, Equations 21 and 22 are the basis for mode-specific molecular
symmetry analysis. One can use the symmetry properties of the Raman polarizability tensor αi j 
and the IR transition dipole μk of a particular mode to analyze the symmetry properties of βq ,i  j  k ,
χq ,i j k , and χq ,eff as defined in Equation 21 (40, 41).
One particularly widely used and tested treatment for the quantitative analysis of SFG-VS
spectra and molecular structure is to derive the βq ,i  j  k tensors of molecular groups of different
symmetry with the empirical bond polarizability model or bond additivity model (38, 39, 41, 107,
110). The molecular symmetry-based formulations of βq ,i  j  k and χq ,i j k have been worked out in
specific detail and are summarized in the literature (40, 41, 107, 125). We do not cover the details
here due to length limitations. However, some of these works need to be constantly consulted if
the quantitative analysis of SFG-VS spectra, both achiral and chiral, is needed (34, 40, 41, 71).
Moreover, with quantum computation, one can directly calculate the mode-specific Raman
polarization tensor and transition dipole values. The parameter from quantum computation can
be plugged in to Equation 22 to help simulate and predict the SFG-VS spectra, in combination
with a molecular dynamics simulation of the interface structure or statistical orientation average
(126–130).

4.6. Polarization Measurement in SFG-VS


(2)
The overall expression for χeff presented in Equations 11a,b can be rearranged to make different
(2)
SFG-VS measurements and for analysis. χeff can usually be written as a linear combination of
experimentally measureable terms corresponding to the seven possible polarization combinations
as follows:
(2)
χeff = [sin cos 1 · χsps + cos sin 1 · χpss + cos cos 1 · χpps ] sin 2
+ [sin sin 1 · χssp + sin cos 1 · χspp + cos sin 1 · χpsp (23a)
+ cos cos 1 · χppp ] cos 2 ,

(2)
χeff,achiral = [sin cos 1 · χsps + cos sin 1 · χpss ] sin 2
(23b)
+ [sin sin 1 · χssp + cos cos 1 · χppp ] cos 2 ,

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 203


PC66CH09-Wang ARI 28 February 2015 12:19

(2)
χeff,chiral = cos cos 1 sin 2 · χpps
(23c)
+ [sin cos 1 · χspp + cos sin 1 · χpsp ] cos 2 ,
with the achiral polarization combination terms

χssp = L y y (ω)L y y (ω1 )L zz (ω2 ) sin β2 · χ y y z ,


χsps = L y y (ω)L zz (ω1 )L y y (ω2 ) sin β1 · χ y zy ,
χpss = L zz (ω)L y y (ω1 )L y y (ω2 ) sin β · χzy y ,
χppp = −L xx (ω)L xx (ω1 )L zz (ω2 ) cos β cos β1 sin β2 · χxxz (24)
− L xx (ω)L zz (ω1 )L xx (ω2 ) cos β sin β1 cos β2 · χxzx
+ L zz (ω)L xx (ω1 )L xx (ω2 ) sin β cos β1 cos β2 · χzxx
+ L zz (ω)L zz (ω1 )L zz (ω2 ) sin β sin β1 sin β2 · χzzz
and the chiral polarization combination terms
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

χspp = L y y (ω)L zz (ω1 )L xx (ω2 ) sin β1 cos β2 · χ y zx


Access provided by 143.255.218.155 on 05/21/20. For personal use only.

+ L y y (ω)L xx (ω1 )L zz (ω2 ) cos β1 sin β2 · χ y xz ,


χpps = L zz (ω)L xx (ω1 )L y y (ω2 ) sin β cos β1 · χzxy
(25)
− L xx (ω)L zz (ω1 )L y y (ω2 ) cos β sin β1 · χxzy ,
χpsp = L zz (ω)L y y (ω1 )L xx (ω2 ) sin β cos β2 · χzy x
− L xx (ω)L y y (ω1 )L zz (ω2 ) cos β sin β2 · χxy z .
Equations 24 and 25 give the seven terms that can be independently measured in SFG experi-
ments, by specifying the polarizations of the sum-frequency detection, visible, and IR fields. As in
Figure 2, the p polarization ( i = 0◦ ) is defined, as the field vector of the incident lasers, or the
detected signal, is within the incident plane (i.e., plane xz), or the plane consists of the interface
normal and the incident laser beam. The s polarization ( i = 90◦ ) occurs when the field vector is
perpendicular to this incident plane. In the SFG literature, the spectrum of the achiral ssp polar-
ization is usually measured, and in the case of a chiral term, the psp polarization is measured (34).

4.7. Polarization Selection Rules and Spectral Assignment


SFG-VS spectra in polarization combinations other than ssp are also important and useful (see
the sidebar Surfaces for Spectroscopy). For example, by comparing the intensity in the ssp, ppp,
and sps polarizations, one can determine the symmetry of the SFG-VS spectral features (41,
131–133). This forms the basis of the polarization selection rules that can be used for spectral
assignment in SFG-VS and for clarifications and corrections of wrong assignments in IR and
Raman studies in the literature (41, 131, 132, 134, 135). A rigorous derivation of the polarization
selection rules in SFG-VS based on the classification of the symmetric and asymmetric Raman
tensors is also available (136). In many ways, such polarization selection rules are similar to the
polarization selection rules in Raman spectroscopy (137), as the symmetry properties of the
Raman tensors significantly contribute to the polarization dependence of the SFG tensors, as
in Equation 22. One successful application of the SFG polarization selection rules clarified the
assignments regarding the asymmetric stretching modes and the symmetric Fermi resonance
vibrational modes of the hydrocarbon methyl and methylene groups (131, 132, 134). Such
fundamental knowledge has begun to impact fields such as biomolecular spectroscopic imaging
(138). One important feature of SFG-VS, which can impact vibrational spectroscopy studies in
general, is the use of the surface or interface to measure and interpret vibrational spectra with
the clearly defined polarization measurements and the polarization selection rules of SFG-VS.

204 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

SURFACES FOR SPECTROSCOPY

SFG-VS has been known as a spectroscopic tool for surface and interface studies. This is the well-accepted idea of
using spectroscopy for surface studies. However, an exciting new possibility of SFG-VS is that we can use surfaces
to study molecular spectroscopy. Especially given its ability to obtain intrinsic spectral lineshapes using HR-BB-
SFG-VS, SFG-VS is ready to play important roles in understanding molecular vibrational spectroscopy. This is
the novel idea of using surface for spectroscopy studies. As it is known, the IR or Raman spectra of large molecules
are complicated and difficult to analyze. SFG-VS offers possible solutions for understanding these spectra. First,
the spectroscopic phase relationship provides SFG-VS with unique and additional spectral resolving power on
overlapping peaks (79, 139, 140). Second, there is a unique set of polarization selection rules in SFG-VS that can
help with spectral assignment (41, 131, 132). Third, as the molecules are partially aligned on the surface, the spectral
lineshape and features are usually less broadened and simpler than those of IR and Raman spectra in the solution or
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

solid phase (73). Finally, because SFG-VS has submonolayer sensitivity, only a tiny amount of chemicals is needed
for an SFG-VS measurement.
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

4.8. Spectra Measured with Different Visible Incident Angles


One usually neglected dimension in SFG-VS measurements involves spectra with different inci-
dent angles. According to Equation 24, χppp is the only χeff term as a linear combination of more
than one χi j k term (34). Because the visible spectrum is with a narrow single frequency, changing
the visible incident angle should not cause a spectral change for all the χeff terms consisting of a
single χi j k term. Therefore, the spectral lineshape changes in the single χi j k term measurement
can be used to check the consistency and reproducibility of the SFG-VS spectra (99).
Conversely, for ppp polarization, changing the visible incident angle would change the spec-
tral interference because the Fresnel factors for each χi j k term are changed differently. Thus, a
significant change in the ppp spectra with a visible incident angle can provide additional spectral
resolving ability for SFG-VS analysis, as one can control the incident angle and tune the interfer-
ence in the ppp spectra to extract additional phase and amplitude information for different spectral
peaks, and even for completely overlapping peaks (41, 79, 133, 139–141). One striking example is
that the intensity of the 3,700-cm−1 free OH peak of the air/neat water interface increased by two
orders of magnitude when the visible incident angle changed from 39◦ to 63◦ from the interface
normal (133), providing an accurate measurement of the water structure at the topmost layer of
the air/water interface. Such interference is unique to second-order nonlinear spectroscopy and
provides additional spectral resolving power to SFG-VS (79).
The power of exploring the interference of different χi j k terms with changing visible incident
angles does not need to be limited to χppp . From Equations 23a–c, one can easily choose a measure-
ment with different combinations of χeff or χi j k with different combinations of polarization angles.
By changing either the visible polarization angle or incident angle, one can tune the interference
between the different χi j k terms, mapping new information from the spectral interference, phase,
and amplitude of different spectral features of complex molecular surfaces and interfaces (139,
140, 142).
Another application of incident angle–dependent SFG-VS measurements is the total inter-
nal reflection (TIR) for solid/air, solid/liquid, and solid/liquid interfaces (19), with which the
SFG-VS signal intensity can be improved by up to two orders of magnitude, as the Fresnel factors
in Equations 12a–c become larger under the TIR condition. Besides improving measurement
sensitivity, TIR-SFG can also avoid laser damage to the materials in the lower phase as the laser
light does not penetrate beyond the diffraction length. However, analysis of TIR-SFG data for

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 205


PC66CH09-Wang ARI 28 February 2015 12:19

spectral interference between different χi j k terms, as in ppp or other mixed-term measurements,


can be difficult because of the complex phase of the Fresnel factors under the TIR condition.

4.9. Polarization Measurement Techniques


Equations 23a–c offer formulations to make measurements with mixed polarizations. This can
be achieved by using polarization angle values other than 90◦ or 0◦ for one or all three of the
polarization angles ( i ).

4.9.1. Polarization null angle technique. The polarization null angle (PNA) technique can
be used to accurately measure the relative amplitude ratio of SFG-VS in different polarizations
and determine the molecular orientation parameter. For an achiral interface or a spectral peak
that does not have a chiral contribution, the PNA experiment is done with a fixed IR beam at
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

p polarization ( 2 = 0◦ ) and the visible beam at polarization 1 = −45◦ (facing the incoming
visible beam and rotating the polarization 45◦ counterclockwise from the p polarization), and the
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

(2)
SFG signal |χeff,achiral |2 is measured as a function of . Then a PNA null that makes the SFG
(2)
signal |χeff,achiral |2 = 0 can be accurately determined from Equation 23b,

0 = χeff,achiral
null
= sin null sin(−45◦ ) · χssp + cos null cos(−45◦ ) · χppp . (26)

Then, one has


χppp
tan null = . (27)
χssp
As null can be accurately obtained by fitting the χeff,achiral ( ) curve against , the χppp /χssp ratio
can be accurately measured. This ratio can be used to determine the orientational parameter as
D =
cos θ /
cos3 θ (41, 143). In a recent example of the PNA technique, investigators uniquely
determined the tilt angles of –CN groups in different Langmuir monolayers with significantly
different phase behaviors (144).

4.9.2. χp(±m)p technique for chiral SFG-VS. χp(±m)p measurements are useful to determine the
chiral spectral response through phase-resolved interference spectra of the achiral ppp and chiral
psp terms (71, 145, 146). Similarly, χs (±m)p measurements can also be made for the achiral ssp and
chiral spp terms (71).
From Equation 23a, by setting 2 = 0◦ , = 0◦ , and 1 = ±m, with m as the intermediate
polarization angles typically at 45◦ , one has

χp(±m)p = cos m · χppp ± sin m · χpsp . (28)

By subtracting the two intensity spectra, one has

Ip(+m)p − Ip(−m)p ∝ 4 cos m sin mRe(χppp χpsp ). (29)

Usually in the literature, one finds that m = 45◦ ; thus, we obtain 4 cos m sin m = 2. In a recent
application of the χp(±m)p and χs(±m)p techniques, researchers studied the intrinsic chirality and
orientational chirality (prochirality) at the chiral R-(+)-limonene and S-(−)-limonene liquid/air
interface (71).

4.9.3. Twin polarization angle method. The twin polarization angle method provides another
mixed polarization measurement for the accurate determination of the weaker chiral terms by
interfering with the stronger achiral terms (116). By using synchronized rotations of both the

206 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

visible and SFG polarization angles, one can use this method to obtain the relative sign and values
of all the achiral and chiral terms, as in the Equation 23a.
Magic angle: a
precisely defined angle
4.10. Orientation Analysis and the SFG-VS Magic Angle from the Legendre
Once the orientational parameter D =
cos θ /
cos3 θ is accurately
 measured, an averaged orien- polynomial expansion
tation angle, or the so-called apparent tilt angle θ = arccos 1/D, can be obtained by assuming of a distribution in
molecular orientation
that the orientational distribution function is very narrow or simply a δ distribution. The validity
angles; for SFG and
of this practice has been questioned after the discovery of the SHG or SFG-VS magic angle (147). SHG, it is
The magic angle of θ = 39.2◦ is obtained with D = 5/3. With the assumption that approximately 39.2◦
the√orientation angle distribution function is a normalized Gaussian distribution, f (θ) =
1/( 2πσ ) exp(−(θ − θ0 )/2σ 2 ), the single D value measured from the SFG polarization experi-
ment is obviously not enough to provide information for two variables (i.e., the center angle θ0
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

and the standard deviation σ ). As illustrated in Figure 3, in the magic angle area, there exists
mathematically an infinitely broad range of (θ0 , σ ) pair values. Therefore, the apparent tilt angle
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

90 90

80 80

70 70

Apparent molecular tilt angle (degrees)


Molecular tilt center angle θ0 (degrees)

60 60

50 50

40 40
SHG magic angle
θ = 35°
30 30
σ = 10°

20 20
σ = 15°

10 10

0 0
0 10 20 30 40 50 60 70 80 90
σ (degrees)
Figure 3
Illustration of the second harmonic generation (SHG) and sum-frequency generation vibrational
spectroscopy (SFG-VS) magic angle and how the physical constraint of the distribution width can
significantly recover the confidence in SFG-VS orientational analysis. The vertical blue line represents the
upper-limit value of σ = 15◦ , the vertical red line represents σ = 10◦ , and the horizontal green line is the
apparent tile angle of the free O–H bond (θ = 35◦ ). Figure adapted from Reference 147. Copyright 1999
American Chemical Society.

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 207


PC66CH09-Wang ARI 28 February 2015 12:19

θ (assuming σ = 0◦ ) can be completely different from the actual tilt angle θ0 with a certain
distribution width [the FWHM of the Gaussian function f (θ) is 2.35σ ].
This magic angle issue, if clearly understood, can improve our understanding of orientation
analysis in SFG and SHG. This is done with a physical constraint on the possible distribution
width σ (Figure 3). One important fact about the molecular interface is that the actual orienta-
tional angle distribution width of a given molecule on the surface or interface has been consistently
shown to have limited values. The best example is the free O–H bond of water molecules at the
topmost layer at the dynamic air/water interface (24, 133). The polarization dependence data
of this narrow spectral feature centered at 3,700 cm−1 suggest that the free O–H bond is quite
ordered at the interface, and analysis of the data put the upper-limit value of σ < 15◦ (133). In the
literature, the apparent tilt angle of the free O–H bond is approximately θ = 35◦ , a value very
close to the magic angle of θ = 39.2◦ . However, with σ < 15◦ , the θ 0 values corresponding to
θ = 35◦ can vary only from 30◦ (σ = 15◦ ) to 35◦ (σ = 0◦ ; i.e., a δ distribution). From Figure 3,
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

one can easily see that with σ < 15◦ , the allowed θ 0 values corresponding to the magic angle
θ = 39.2◦ can vary only less than 3◦ from σ = 15◦ to σ = 0◦ . Therefore, as long as the σ
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

value is limited, the deviation of θ from θ 0 is significantly limited, even in the magic angle
region.
σ = 15◦ is already a fairly broad distribution, as the corresponding FWHM is approximately

35 . In fact, the spectral features observed in SFG-VS measurements are always sensitive to the
measurement polarizations, and most of the liquid or material interfaces are less dynamic than
the air/water interface. Therefore, the orientational distribution width of the molecular groups
can be even narrower than the air/water interface. With such a constraint in the actual physical
world, the deviation of the apparent θ value obtained from the assumption of σ = 0◦ is usually
not much different from the mathematically allowed θ 0 values, even though the exact value of
σ is often not known. Multiple orientation angles of similar groups at an interface do exist, and
they can be resolved through accurately measured orientation parameter D values (148, 149), or
through the subtle spectral splitting of the same molecular groups with slightly different chemical
environments or orientations in different polarizations using HR-BB-SFG-VS (41). Therefore, the
existence of the SHG or SFG-VS magic angle should not affect the main conclusions drawn from
SHG or SFG-VS orientational analysis, provided the orientational parameter D was determined
with enough accuracy.

5. CONCLUDING REMARKS
SFG-VS is a unique in situ spectroscopic tool for molecular surface and interface studies. Although
the application of SFG-VS has broadened, and many complex molecular surfaces and interfaces
have been studied in the past 30 years, the quantitative measurement and analysis of SFG-VS data
are still limited. To obtain detailed structural and dynamics information for molecular surfaces
and interfaces, the measurement and analysis of the detailed spectral lineshape, polarization de-
pendence, and orientation analysis have become increasingly important. The recent development
of sub-wave-number HR-BB-SFG-VS has provided new opportunities in experimentally obtain-
ing SFG-VS spectra with nearly intrinsic lineshapes with minimal broadening and distortion.
This development provided a direct test for a unified theory on the frequency-domain and time-
domain lineshapes, through the accurate fitting of the frequency spectra for spectral and coherent
dynamics parameters of the molecular interface. It also provided ample opportunities for quan-
titative polarization measurements and orientation analysis of SFG-VS spectra. Unprecedented
detailed information on the molecular structure and dynamics can be obtained by employing the
concepts of the intrinsic spectral lineshape, polarization dependence, and molecular orientation

208 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

in SFG-VS measurements and analysis. These concepts can also be extended to other nonlinear
spectroscopies.
As SFG and SHG increase in their popularity among researchers, a clear understanding of the
capabilities and limitations of the techniques is necessary. As described in this review, many appar-
ent past limitations can be attributed to a lack of a proper theoretical or experimental approach.
As the future of the field looks bright for applications in complex interfaces, an understanding of
the proper treatment of the lineshape and polarization analysis becomes of great importance in
making SFG a quantitative and reliable tool.

SUMMARY POINTS
1. SFG-VS from both chiral and achiral surfaces in the reflective geometry is dominated
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

by surface dipole contributions.


2. The SFG-VS spectral lineshape in the frequency domain is affected by the line width
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

of the visible laser profile. SFG with sub-wave-number resolution can provide nearly
intrinsic spectral lineshapes in SFG-VS spectra.
3. Frequency-domain and time-domain SFG-VS spectral lineshapes are equivalent and
have the same spectral and coherent dynamics parameters. High-resolution frequency-
domain measurements are advantageous for obtaining lineshape parameters of complex
spectra.
4. Polarization measurements for specific information regarding achiral and chiral surfaces
and interfaces can be designed through the complete expression of the effective SFG-VS
susceptibility.
5. Polarization and molecular symmetry analysis not only can provide general rules for
vibrational spectra assignment for SFG-VS, but also can help in understanding and
assigning IR and Raman spectra.
6. Knowledge of the physical constraints on the orientational distribution width can break
the spell of the SHG and SFG-VS magic angle.

FUTURE ISSUES
1. Sub-wave-number HR-BB-SFG-VS has shown great potential in obtaining intrinsic
spectral lineshapes of molecular species at interfaces (73, 79, 80). For broader applica-
tions, its sensitivity needs to be improved, and its cost needs to be reduced with better
engineering designs.
2. SFG-VS measurements already have submonolayer sensitivity. However, the SFG-VS
signal sensitivity needs to be significantly improved, ideally by two or three orders of
magnitude, to study molecular groups at relatively small surface coverages (e.g., the
reaction center or vibrationally labeled residues in macromolecules or proteins).
3. SFG-VS has unique potential in biophysical studies, especially those involving mem-
brane processes with proteins, lipids, and polysaccharides (34). Chiral SFG can be ap-
plied to study protein (or DNA) interaction and structural changes during biological
processes.

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 209


PC66CH09-Wang ARI 28 February 2015 12:19

4. Doubly resonant SFG-VS can be used to study both electronic and vibrational spec-
troscopy of molecular interfaces. Most importantly, the coupling between the electronic
and vibrational motion within the molecule at the interface can be probed. There are
many future applications in understanding electron and charge transfer processes at sur-
faces and interfaces. Quantitative measurement and analysis in this direction are still
limited.
5. SFG-VS and SHG scattering, or second-order Mie scattering, can be used to study the
surfaces of colloidal particles in solution or aerosol particles in the atmosphere or gas
phase (7, 29). Quantitative measurement and analysis in this direction can be useful and
exciting.
6. The ability of SFG-VS to probe buried interfaces, such as the solid/liquid interface, is
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

limited by the transmission of the IR pulse. However, the interface selectivity and sub-
monolayer selectivity of SFG-VS still make it a better choice for studying buried inter-
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

faces than other IR techniques. SFG-VS experimental cell design needs to be improved
for the study of buried interfaces, with many applications in catalysis, geochemistry,
environmental chemistry, etc.

DISCLOSURE STATEMENT
The authors are not aware of any affiliations, memberships, funding, or financial holdings that
might be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS
L.V., W.G., and L.F. contributed equally to this review. L.F. is the William R. Wiley Postdoctoral
Fellow at the William R. Wiley Environmental Molecular Sciences Laboratory (EMSL). Part
of this work was supported by the Pacific Northwest National Laboratory (PNNL) and was
conducted at the EMSL, a national scientific user facility located at the PNNL and sponsored by
the Department of Energy’s Office of Biological and Environmental Research.

LITERATURE CITED
1. Eisenthal KB. 1992. Equilibrium and dynamic processes at interfaces by second harmonic and sum
frequency generation. Annu. Rev. Phys. Chem. 43:627–61
2. Richmond GL. 2001. Structure and bonding of molecules at aqueous surfaces. Annu. Rev. Phys. Chem.
52:357–89
3. Chen Z, Shen YR, Somorjai GA. 2002. Studies of polymer surfaces by sum frequency generation vibra-
tional spectroscopy. Annu. Rev. Phys. Chem. 53:437–65
4. Geiger FM. 2009. Second harmonic generation, sum frequency generation, and χ (3) : dissecting envi-
ronmental interfaces with a nonlinear optical Swiss Army knife. Annu. Rev. Phys. Chem. 60:61–83
5. Haupert LM, Simpson GJ. 2009. Chirality in nonlinear optics. Annu. Rev. Phys. Chem. 60:345–65
6. Jubb AM, Hua W, Allen HC. 2012. Environmental chemistry at vapor/water interfaces: insights from
vibrational sum frequency generation spectroscopy. Annu. Rev. Phys. Chem. 63:107–30
7. Roke S, Gonella G. 2012. Nonlinear light scattering and spectroscopy of particles and droplets in liquids.
Annu. Rev. Phys. Chem. 63:353–78
8. Nihonyanagi S, Mondal JA, Yamaguchi S, Tahara T. 2013. Structure and dynamics of interfacial water
studied by heterodyne-detected vibrational sum-frequency generation. Annu. Rev. Phys. Chem. 64:579–
603

210 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

9. Shen YR. 2013. Phase-sensitive sum-frequency spectroscopy. Annu. Rev. Phys. Chem. 64:129–50
10. Shen YR. 1984. The Principles of Nonlinear Optics. New York: Wiley Intersci.
11. Shen Y. 2012. Basic theory of surface sum-frequency generation. J. Phys. Chem. C 116:15505–9
12. Shen YR. 1989. Surface properties probed by second-harmonic and sum-frequency generation. Nature
337:519–25
13. Shen YR. 2000. Surface nonlinear optics: a historical perspective. IEEE J. Sel. Top. Quant. Electron.
6:1375–79
14. Chen CK, de Castro ARB, Shen YR. 1981. Surface-enhanced second-harmonic generation. Phys. Rev.
Lett. 46:145–48
15. Chen CK, Heinz TF, Ricard D, Shen YR. 1981. Detection of molecular monolayers by optical second-
harmonic generation. Phys. Rev. Lett. 46:1010–12
16. Heinz TF, Chen CK, Ricard D, Shen YR. 1981. Optical second-harmonic generation from a monolayer
of centrosymmetric molecules adsorbed on silver. Chem. Phys. Lett. 83:180–82
17. Tom HWK. 1984. Studies of surfaces using optical second-harmonic generation. PhD Thesis, Univ. Calif.,
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

Berkeley
18. Zhu XD, Suhr H, Shen YR. 1987. Surface vibrational spectroscopy by infrared-visible sum frequency
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

generation. Phys. Rev. B 35:3047–50


19. Buck M, Himmelhaus M. 2001. Vibrational spectroscopy of interfaces by infrared-visible sum frequency
generation. J. Vac. Sci. Technol. A 19:2717–36
20. Hicks JM, Kemnitz K, Eisenthal KB. 1986. Studies of liquid surfaces by second harmonic generation.
J. Phys. Chem. 90:560–62
21. Eisenthal KB. 1993. Liquid interfaces. Acc. Chem. Res. 26:636–43
22. Miranda PB, Shen YR. 1999. Liquid interfaces: a study by sum-frequency vibrational spectroscopy.
J. Phys. Chem. B 103:3292–307
23. Eisenthal KB. 1996. Liquid interfaces probed by second-harmonic and sum-frequency spectroscopy.
Chem. Rev. 96:1343–60
24. Du Q, Superfine R, Freysz E, Shen YR. 1993. Vibrational spectroscopy of water at the vapor water
interface. Phys. Rev. Lett. 70:2313–16
25. Scatena LF, Brown MG, Richmond GL. 2001. Water at hydrophobic surfaces: weak hydrogen bonding
and strong orientation effects. Science 292:908–12
26. Richmond GL. 2002. Molecular bonding and interactions at aqueous surfaces as probed by vibrational
sum frequency spectroscopy. Chem. Rev. 102:2693–724
27. Shen YR, Ostroverkhov V. 2006. Sum-frequency vibrational spectroscopy on water interfaces: polar
orientation of water molecules at interfaces. Chem. Rev. 106:1140–54
28. Wang H, Yan ECY, Borguet E, Eisenthal KB. 1996. Second harmonic generation from the surface of
centrosymmetric particles in bulk solution. Chem. Phys. Lett. 259:15–20
29. Eisenthal KB. 2006. Second harmonic spectroscopy of aqueous nano- and microparticle interfaces. Chem.
Rev. 106:1462–77
30. Somorjai GA, Rupprechter G. 1999. Molecular studies of catalytic reactions on crystal surfaces at high
pressures and high temperatures by infrared-visible sum frequency generation (SFG) surface vibrational
spectroscopy. J. Phys. Chem. B 103:1623–38
31. Somorjai GA, McCrea KR. 2000. Sum frequency generation: surface vibrational spectroscopy studies of
catalytic reactions on metal single-crystal surfaces. Adv. Catal. 45:385–438
32. Chen Z. 2010. Investigating buried polymer interfaces using sum frequency generation vibrational spec-
troscopy. Prog. Polym. Sci. 35:1376–402
33. Chen XY, Clarke ML, Wang J, Chen Z. 2005. Sum frequency generation vibrational spectroscopy 34. Up-to-date review
studies on molecular conformation and orientation of biological molecules at interfaces. Int. J. Mod. on chiral SFG studies
Phys. B 19:691–713 on biological
34. Yan ECY, Fu L, Wang Z, Liu W. 2014. Biological macromolecules at interfaces probed by chiral macromolecules at
vibrational sum frequency generation spectroscopy. Chem. Rev. 114:8471–98 interfaces.
35. Roy S, Covert PA, FitzGerald WR, Hore DK. 2014. Biomolecular structure at solid-liquid interfaces as
revealed by nonlinear optical spectroscopy. Chem. Rev. 114:8388–415

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 211


PC66CH09-Wang ARI 28 February 2015 12:19

36. Smith JP, Hinson-Smith V. 2004. SFG coming of age. Anal. Chem. 76:287A–90A
38. Detailed description 37. Richter LJ, Petralli-Mallow TP, Stephenson JC. 1998. Vibrationally resolved sum-frequency generation
of quantitative with broad-bandwidth infrared pulses. Opt. Lett. 23:1594–96
polarization and 38. Zhuang X, Miranda PB, Kim D, Shen YR. 1999. Mapping molecular orientation and conforma-
orientation analysis with tion at interfaces by surface nonlinear optics. Phys. Rev. B 59:12632–40
SHG and SFG-VS.
39. Wei X, Hong SC, Zhuang XW, Goto T, Shen YR. 2000. Nonlinear optical studies of liquid
crystal alignment on a rubbed polyvinyl alcohol surface. Phys. Rev. E 62:5160–72
39. Clarification of 40. Moad AJ, Simpson GJ. 2004. A unified treatment of selection rules and symmetry relations for
Fresnel and local field sum-frequency and second harmonic spectroscopies. J. Phys. Chem. B 108:3548–62
factors in SFG-VS
41. Wang HF, Gan W, Lu R, Rao Y, Wu BH. 2005. Quantitative spectral and orientational analysis
analysis.
in surface sum frequency generation vibrational spectroscopy (SFG-VS). Int. Rev. Phys. Chem.
40. Comprehensive 24:191–256
derivation of 42. Zheng DS, Wang Y, Liu AA, Wang HF. 2008. Microscopic molecular optics theory of surface second
macroscopic and harmonic generation and sum-frequency generation spectroscopy based on the discrete dipole lattice
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

microscopic molecular model. Int. Rev. Phys. Chem. 27:629–64


responses in SFG-VS 43. Shen Y. 1994. Surface spectroscopy by nonlinear optics. In Frontiers in Laser Spectroscopy: Proc. Int. Sch.
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

and SHG.
Phys. “Enrico Fermi” Course CXX, ed. TW Hansch, M Inguscio, pp. 139–65. Amsterdam: North-Holland
41. Formulations and 44. Andrews DL, Blake NP. 1988. Forbidden nature of multipolar contributions to second-harmonic gen-
examples of quantitative eration in isotropic fluids. Phys. Rev. A 38:3113–15
polarization 45. Heinz TF, DiVincenzo DP. 1990. Comment on “Forbidden nature of multipolar contributions to
measurement and second-harmonic generation in isotropic fluids.” Phys. Rev. A 42:6249–51
analysis. 46. Zhu XD, Shen YR. 1990. Multipolar contributions to optical second-harmonic generation in isotropic
fluids. Phys. Rev. A 41:4549
47. Shen YR. 1999. Surface contribution versus bulk contribution in surface nonlinear optical spectroscopy.
Appl. Phys. B 68:295–300
48. Wei X, Hong SC, Lvovsky AI, Held H, Shen YR. 2000. Evaluation of surface versus bulk contributions
in sum-frequency vibrational spectroscopy using reflection and transmission geometries. J. Phys. Chem.
B 104:3349–54
49. Held H, Lvovsky AI, Wei X, Shen YR. 2002. Bulk contribution from isotropic media in surface sum-
frequency generation. Phys. Rev. B 66:205110
50. Morita A. 2004. Toward computation of bulk quadrupolar signals in vibrational sum frequency generation
spectroscopy. Chem. Phys. Lett. 398:361–66
51. Goh MC, Hicks JM, Kemnitz K, Pinto GR, Heinz TF, et al. 1988. Absolute orientation of water
molecules at the neat water surface. J. Phys. Chem. 92:5074–75
52. Zhang WK, Zheng DS, Xu YY, Bian HT, Guo Y, Wang HF. 2005. Reconsideration of second-harmonic
generation from isotropic liquid interface: broken Kleinman symmetry of neat air/water interface from
dipolar contribution. J. Chem. Phys. 123:224713
53. Franken PA, Ward JF. 1963. Optical harmonics and nonlinear phenomena. Rev. Mod. Phys. 35:23–39
54. Dailey CA, Burke BJ, Simpson GJ. 2004. The general failure of Kleinman symmetry in practical nonlinear
optical applications. Chem. Phys. Lett. 390:8–13
55. Martin-Gassin G, Benichou E, Bachelier G, Russier-Antoine I, Jonin C, Brevet PF. 2008. Compres-
sion induced chirality in dense molecular films at the air-water interface probed by second harmonic
generation. J. Phys. Chem. C 112:12958–65
56. Xu Y-Y, Wei F, Wang H-F. 2009. Comment on “Compression induced chirality in dense molecular
films at the air-water interface probed by second harmonic generation.” J. Phys. Chem. C 113:4222–26
57. Rodrı́guez FJ, Wang FX, Canfield BK, Cattaneo S, Kauranen M. 2007. Multipolar tensor analysis of
second-order nonlinear optical response of surface and bulk of glass. Opt. Express 15:8695–701
58. Hommel EL, Allen HC. 2003. The air-liquid interface of benzene, toluene, m-xylene, and mesitylene:
a sum frequency, Raman, and infrared spectroscopic study. Analyst 128:750–55
59. Kawaguchi T, Shiratori K, Henmi Y, Ishiyama T, Morita A. 2012. Mechanisms of sum frequency
generation from liquid benzene: symmetry breaking at interface and bulk contribution. J. Phys. Chem. C
116:13169–82

212 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

60. Giordmaine JA. 1965. Nonlinear optical properties of liquids. Phys. Rev. 138:A1599–606
61. Rentzepis PM, Giordmaine JA, Wecht KW. 1966. Coherent optical mixing in optically active liquids.
Phys. Rev. Lett. 16:792–94
62. Ji N, Shen YR. 2006. A novel spectroscopic probe for molecular chirality. Chirality 18:146–58
63. Belkin MA, Han SH, Wei X, Shen YR. 2001. Sum-frequency generation in chiral liquids near electronic
resonance. Phys. Rev. Lett. 87:113001
64. Fischer P, Wiersma DS, Righini R, Champagne B, Buckingham AD. 2000. Three-wave mixing in chiral
liquids. Phys. Rev. Lett. 85:4253–56
65. Belkin MA, Shen YR, Harris RA. 2004. Sum-frequency vibrational spectroscopy of chiral liquids off and
close to electronic resonance and the antisymmetric Raman tensor. J. Chem. Phys. 120:10118–26
66. Belkin MA, Shen YR. 2005. Non-linear optical spectroscopy as a novel probe for molecular chirality.
Int. Rev. Phys. Chem. 24:257–99
67. Belkin MA, Shen YR. 2003. Doubly resonant IR-UV sum-frequency vibrational spectroscopy on molec-
ular chirality. Phys. Rev. Lett. 91:213907
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

68. Han SH, Ji N, Belkin MA, Shen YR. 2002. Sum-frequency spectroscopy of electronic resonances on a
chiral surface monolayer of bi-naphthol. Phys. Rev. B 66:165415
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

69. Wang J, Chen XY, Clarke ML, Chen Z. 2005. Detection of chiral sum frequency generation vibrational
spectra of proteins and peptides at interfaces in situ. Proc. Natl. Acad. Sci. USA 102:4978–83
70. Champagne B, Fischer P, Buckingham AD. 2000. Ab initio investigation of the sum-frequency hyper-
polarizability of small chiral molecules. Chem. Phys. Lett. 331:83–88
71. Fu L, Zhang Y, Wei Z-H, Wang H-F. 2014. Intrinsic chirality and prochirality at air/R-(+)- and S- 71. Detailed
(−)-limonene interfaces: spectral signatures with interference chiral sum-frequency generation demonstration of chiral
vibrational spectroscopy. Chirality 26:509–20 SFG-VS on intrinsic
72. Owrutsky JC, Culver JP, Li M, Kim YR, Sarisky MJ, et al. 1992. Femtosecond coherent transient infrared chirality and
spectroscopy of CO on Cu(111). J. Chem. Phys. 97:4421–27 prochirality.
73. Velarde L, Wang HF. 2013. Unified treatment and measurement of the spectral resolution
and temporal effects in frequency-resolved sum-frequency generation vibrational spectroscopy 73. Unified treatment
(SFG-VS). Phys. Chem. Chem. Phys. 15:19970–84 on frequency- and
time-domain SFG-VS
74. Laaser JE, Xiong W, Zanni MT. 2011. Time-domain SFG spectroscopy using mid-IR pulse
with HR-BB-SFG-VS
shaping: practical and intrinsic advantages. J. Phys. Chem. B 115:2536–46
experimental validation.
75. McGuire JA, Shen YR. 2006. Signal and noise in Fourier-transform sum-frequency surface vibrational
spectroscopy with femtosecond lasers. J. Opt. Soc. Am. B 23:363–69
76. Guyot-Sionnest P. 1991. Coherent processes at surfaces: free-induction decay and photon echo of the 74. Heterodyne
Si-H stretching vibration for H/Si(111). Phys. Rev. Lett. 66:1489–92 SFG-VS in the time
domain pushed to its
77. Bordenyuk AN, Jayathilake H, Benderskii AV. 2005. Coherent vibrational quantum beats as a probe of
limit.
Langmuir-Blodgett monolayers. J. Phys. Chem. B 109:15941–49
78. Roke S, Kleyn AW, Bonn M. 2003. Time- versus frequency-domain femtosecond surface sum frequency
generation. Chem. Phys. Lett. 370:227–32
79. Velarde L, Zhang XY, Lu Z, Joly AG, Wang ZM, Wang HF. 2011. Communication: Spectroscopic 79. First report on
phase and lineshapes in high-resolution broadband sum frequency vibrational spectroscopy: HR-BB-SFG-VS,
resolving interfacial inhomogeneities of “identical” molecular groups. J. Chem. Phys. 135: demonstrating the
241102 unique spectral
80. Velarde L, Wang HF. 2013. Capturing inhomogeneous broadening of the –CN stretch vibration resolving capability of
SFG-VS.
in a Langmuir monolayer with high-resolution spectra and ultrafast vibrational dynamics in
sum-frequency generation vibrational spectroscopy (SFG-VS). J. Chem. Phys. 139:084204
81. Shen YR. 1998. Sum frequency generation (SFG) spectroscopy. In Nonlinear Spectroscopy for Molecular 80. Intrinsic SFG-VS
Structure Determination, ed. RW Field, E Hirota, JP Maier, pp. 249–71. Oxford, UK: Blackwell Sci. spectral lineshape
82. Nihonyanagi S, Eftekhari-Bafrooei A, Borguet E. 2011. Ultrafast vibrational dynamics and spectroscopy analysis capturing
of a siloxane self-assembled monolayer. J. Chem. Phys. 134:084701 inhomogeneous
lineshapes.
83. Olivero JJ, Longbothum RL. 1977. Empirical fits to the Voigt line width: a brief review. J. Quant.
Spectrosc. Radiat. Transfer 17:233–36
84. Sovago M, Vartiainen E, Bonn M. 2009. Determining absolute molecular orientation at interfaces: a
phase retrieval approach for sum frequency generation spectroscopy. J. Phys. Chem. C 113:6100–6

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 213


PC66CH09-Wang ARI 28 February 2015 12:19

85. Weeraman C, Mitchell SA, Lausten R, Johnston LJ, Stolow A. 2010. Vibrational sum frequency gener-
ation spectroscopy using inverted visible pulses. Opt. Express 18:11483–94
86. Stiopkin IV, Jayathilake HD, Weeraman C, Benderskii AV. 2010. Temporal effects on spectroscopic
line shapes, resolution, and sensitivity of the broad-band sum frequency generation. J. Chem. Phys.
132:234503
87. Carter JA, Wang ZH, Dlott DD. 2009. Ultrafast nonlinear coherent vibrational sum-frequency spec-
troscopy methods to study thermal conductance of molecules at interfaces. Acc. Chem. Res. 42:1343–51
88. Shalhout FY, Malyk S, Benderskii AV. 2012. Relative phase change of nearby resonances in temporally
delayed sum frequency spectra. J. Phys. Chem. Lett. 3:3493–97
89. Mukamel S. 1995. Principles of Nonlinear Optical Spectroscopy. New York: Oxford Univ. Press
90. Hamm P, Zanni M. 2011. Concepts and Methods of 2D Infrared Spectroscopy. Cambridge, UK: Cambridge
Univ. Press
91. Verreault D, Hua W, Allen HC. 2012. From conventional to phase-sensitive vibrational sum frequency
generation spectroscopy: probing water organization at aqueous interfaces. J. Phys. Chem. Lett. 3:3012–28
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

92. Ostroverkhov V, Waychunas GA, Shen YR. 2005. New information on water interfacial structure re-
vealed by phase-sensitive surface spectroscopy. Phys. Rev. Lett. 94:046102
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

93. Ji N, Ostroverkhov V, Tian CS, Shen YR. 2008. Characterization of vibrational resonances of water-
vapor interfaces by phase-sensitive sum-frequency spectroscopy. Phys. Rev. Lett. 100:096102
94. Tian CS, Shen YR. 2009. Sum-frequency vibrational spectroscopic studies of water/vapor interfaces.
Chem. Phys. Lett. 470:1–6
95. Nihonyanagi S, Yamaguchi S, Tahara T. 2009. Direct evidence for orientational flip-flop of water
molecules at charged interfaces: a heterodyne-detected vibrational sum frequency generation study.
J. Chem. Phys. 130:204704
96. Mondal JA, Nihonyanagi S, Yamaguchi S, Tahara T. 2010. Structure and orientation of water at charged
lipid monolayer/water interfaces probed by heterodyne-detected vibrational sum frequency generation
spectroscopy. J. Am. Chem. Soc. 132:10656–57
97. Mondal JA, Nihonyanagi S, Yamaguchi S, Tahara T. 2012. Three distinct water structures at a zwit-
terionic lipid/water interface revealed by heterodyne-detected vibrational sum frequency generation.
J. Am. Chem. Soc. 134:7842–50
98. Hua W, Chen XK, Allen HC. 2011. Phase-sensitive sum frequency revealing accommodation of bicar-
bonate ions, and charge separation of sodium and carbonate ions within the air/water interface. J. Phys.
Chem. A 115:6233–38
99. Feng RR, Guo Y, Lu R, Velarde L, Wang HF. 2011. Consistency in the sum frequency generation
intensity and phase vibrational spectra of the air/neat water interface. J. Phys. Chem. A 115:6015–27
100. Pool RE, Versluis J, Backus EHG, Bonn M. 2011. Comparative study of direct and phase-specific vibra-
tional sum-frequency generation spectroscopy: advantages and limitations. J. Phys. Chem. B 115:15362–
69
101. Yang PK, Huang JY. 1997. Phase-retrieval problems in infrared-visible sum-frequency generation spec-
troscopy by the maximum-entropy method. J. Opt. Soc. Am. B 14:2443–48
102. de Beer AGF, Samson JS, Hua W, Huang ZS, Chen XK, et al. 2011. Direct comparison of phase-sensitive
vibrational sum frequency generation with maximum entropy method: case study of water. J. Chem. Phys.
135:224701
103. Nihonyanagi S, Singh PC, Yamaguchi S, Tahara T. 2012. Ultrafast vibrational dynamics of a charged
aqueous interface by femtosecond time-resolved heterodyne-detected vibrational sum frequency gener-
ation. Bull. Chem. Soc. Jpn. 85:758–60
104. Singh PC, Nihonyanagi S, Yamaguchi S, Tahara T. 2012. Ultrafast vibrational dynamics of water at a
charged interface revealed by two-dimensional heterodyne-detected vibrational sum frequency genera-
tion. J. Chem. Phys. 137:094706
105. Miranda PB. 1998. Nonlinear vibrational spectroscopy of surfactants at liquid interfaces. PhD Thesis, Univ.
Calif., Berkeley
106. Wei X. 2000. Sum-frequency spectroscopic studies: I. Surface melting of ice. II. Surface alignment of polymers.
PhD Thesis, Univ. Calif., Berkeley

214 Wang et al.


PC66CH09-Wang ARI 28 February 2015 12:19

107. Hirose C, Akamatsu N, Domen K. 1992. Formulas for the analysis of surface sum-frequency generation
spectrum by CH stretching modes of methyl and methylene groups. J. Chem. Phys. 96:997–1004
108. Bain CD. 1995. Sum-frequency vibrational spectroscopy of the solid-liquid interface. J. Chem. Soc.
Faraday Trans. 91:1281–96
109. Zhang D, Gutow J, Eisenthal KB. 1994. Vibrational spectra, orientations, and phase transitions in
long-chain amphiphiles at the air-water interface: probing the head and tail groups by sum-frequency
generation. J. Phys. Chem. 98:13729–34
110. Hirose C, Yamamoto H, Akamatsu N, Domen K. 1993. Orientation analysis by simulation of vibrational
sum-frequency generation spectrum: CH stretching bands of the methyl group. J. Phys. Chem. 97:10064–
69
111. Hirose C, Akamatsu N, Domen K. 1992. Formulas for the analysis of the surface SFG spectrum and
transformation coefficients of cartesian SFG tensor components. Appl. Spectrosc. 46:1051–72
112. Akamatsu N, Domen K, Hirose C. 1993. SFG study of two-dimensional orientation of surface methyl
groups on cadmium arachidate Langmuir-Blodgett-films. J. Phys. Chem. 97:10070–75
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

113. Simpson GJ. 2004. Molecular origins of the remarkable chiral sensitivity of second-order nonlinear
optics. ChemPhysChem 5:1301–10
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

114. Bloembergen N, Pershan PS. 1962. Light waves at the boundary of nonlinear media. Phys. Rev. 128:606–
22
115. Wang H-F. 2012. In situ measurement of chirality of molecules and molecular assemblies with surface
nonlinear spectroscopy. In Comprehensive Chiroptical Spectroscopy, Vol. 1: Instrumentation, Methodologies,
and Theoretical Simulations, ed. N Berova, PL Polavarapu, K Nakanishi, RW Woody, pp. 373–406. New
York: Wiley
116. Wei F, Xu Y-Y, Guo Y, Liu S-L, Wang H-F. 2009. Quantitative surface chirality detection with sum
frequency generation vibrational spectroscopy: twin polarization angle approach. Chin. J. Chem. Phys.
22:592–600
117. Ye PX, Shen YR. 1983. Local-field effect on linear and non-linear optical-properties of adsorbed
molecules. Phys. Rev. B 28:4288–94
118. Goldstein H. 2000. Classical Mechanics. New York: Addison-Wesley
119. Michl J, Thulstrup EW. 1986. Spectroscopy with Polarized Light: Solute Alignment by Photoselection in Liquid
Crystals, Polymers, and Membranes. Weinheim: VCH
120. Roy S, Hung K-K, Stege U, Hore DK. 2013. Rotations, projections, direction cosines, and vibrational
spectra. Appl. Spectrosc. Rev. 49:233–48
121. Rao Y, Tao YS, Wang HF. 2003. Quantitative analysis of orientational order in the molecular monolayer
by surface second harmonic generation. J. Chem. Phys. 119:5226–36
122. Du Q, Freysz E, Shen YR. 1994. Surface vibrational spectroscopic studies of hydrogen bonding and
hydrophobicity. Science 264:826–28
123. Superfine R, Huang JY, Shen YR. 1990. Experimental determination of the sign of molecular dipole
moment derivatives: an infrared-visible sum frequency generation absolute phase measurement study.
Chem. Phys. Lett. 172:303–6
124. Armstrong JA, Bloembergen N, Ducuing J, Pershan PS. 1962. Interactions between light waves in a
nonlinear dielectric. Phys. Rev. 127:1918–39
125. Wu H, Zhang WK, Gan W, Cui ZF, Wang HF. 2006. An empirical approach to the bond additivity
model in quantitative interpretation of sum frequency generation vibrational spectra. J. Chem. Phys.
125:133203
126. Hore DK, Beaman DK, Parks DH, Richmond GL. 2005. Whole-molecule approach for determining
orientation at isotropic surfaces by nonlinear vibrational spectroscopy. J. Phys. Chem. B 109:16846–51
127. Yeh YL, Zhang C, Held H, Mebel AM, Wei X, et al. 2001. Structure of the acetone liquid/vapor interface.
J. Chem. Phys. 114:1837–43
128. Hall SA, Jena KC, Covert PA, Roy S, Trudeau TG, Hore DK. 2014. Molecular-level surface structure
from nonlinear vibrational spectroscopy combined with simulations. J. Phys. Chem. B 118:5617–36
129. Ishiyama T, Imamura T, Morita A. 2014. Theoretical studies of structures and vibrational sum frequency
generation spectra at aqueous interfaces. Chem. Rev. 114:8447–70

www.annualreviews.org • SFG-VS: Lineshape, Polarization, and Orientation 215


PC66CH09-Wang ARI 28 February 2015 12:19

130. Morita A, Ishiyama T. 2008. Recent progress in theoretical analysis of vibrational sum frequency gen-
eration spectroscopy. Phys. Chem. Chem. Phys. 10:5801–16
131. Lu R, Gan W, Wu BH, Chen H, Wang HF. 2004. Vibrational polarization spectroscopy of CH stretching
modes of the methylene group at the vapor/liquid interfaces with sum frequency generation. J. Phys.
Chem. B 108:7297–306
132. Lu R, Gan W, Wu BH, Zhang Z, Guo Y, Wang HF. 2005. C–H stretching vibrations of methyl,
methylene and methine groups at the vapor/alcohol (n = 1–8) interfaces. J. Phys. Chem. B 109:14118–29
133. Gan W, Wu D, Zhang Z, Feng RR, Wang HF. 2006. Polarization and experimental configuration
analyses of sum frequency generation vibrational spectra, structure, and orientational motion of the
air/water interface. J. Chem. Phys. 124:114705
134. Gan W, Zhang Z, Feng RR, Wang HF. 2006. Identification of overlapping features in the sum frequency
generation vibrational spectra of air/ethanol interface. Chem. Phys. Lett. 423:261–65
135. Yu YQ, Lin K, Zhou XG, Wang H, Liu SL, Ma XX. 2007. New C–H stretching vibrational spectral
features in the Raman spectra of gaseous and liquid ethanol. J. Phys. Chem. C 111:8971–78
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

136. Rivera CA, Fourkas JT. 2011. Reexamining the interpretation of vibrational sum-frequency generation
spectra. Int. Rev. Phys. Chem. 30:409–43
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

137. Long DA. 2002. The Raman Effect: A Unified Treatment of the Theory of Raman Scattering by Molecules.
New York: Wiley
138. Chung CY, Boik J, Potma EO. 2013. Biomolecular imaging with coherent nonlinear vibrational mi-
croscopy. Annu. Rev. Phys. Chem. 64:77–99
139. Gan W, Wu BH, Zhang Z, Guo Y, Wang HF. 2007. Vibrational spectra and molecular orientation
with experimental configuration analysis in surface sum frequency generation (SFG). J. Phys. Chem. C
111:8716–25
140. Gan W, Zhang Z, Feng RR, Wang HF. 2007. Spectral interference and molecular conformation at
liquid interface with sum frequency generation vibrational spectroscopy (SFG-VS). J. Phys. Chem. C
111:8726–38
141. Zhang Z, Guo Y, Lu Z, Velarde L, Wang HF. 2012. Resolving two closely overlapping –CN vibrations
and structure in the Langmuir monolayer of the long-chain nonadecanenitrile by polarization sum
frequency generation vibrational spectroscopy. J. Phys. Chem. C 116:2976–87
142. Wang J, Clarke ML, Chen Z. 2004. Polarization mapping: a method to improve sum frequency gener-
ation spectral analysis. Anal. Chem. 76:2159–67
143. Lu R, Gan W, Wang HF. 2003. Novel method for accurate determination of the orientational angle of
interfacial chemical groups. Chin. Sci. Bull. 48:2183–87
144. Velarde L, Wang HF. 2013. Unique determination of the –CN group tilt angle in Langmuir monolayers
using sum-frequency polarization null angle and phase. Chem. Phys. Lett. 585:42–48
145. Stokes GY, Gibbs-Davis JM, Boman FC, Stepp BR, Condie AG, et al. 2007. Making “sense” of DNA.
J. Am. Chem. Soc. 129:7492–93
146. Belkin MA, Kulakov TA, Ernst KH, Yan L, Shen YR. 2000. Sum-frequency vibrational spectroscopy on
chiral liquids: a novel technique to probe molecular chirality. Phys. Rev. Lett. 85:4474–77
147. Simpson GJ, Rowlen KL. 1999. An SHG magic angle: dependence of second harmonic generation
orientation measurements on the width of the orientation distribution. J. Am. Chem. Soc. 121:2635–36
148. Chen H, Gan W, Wu BH, Wu D, Zhang Z, Wang HF. 2005. Determination of the two methyl group
orientations at vapor/acetone interface with polarization null angle method in SFG vibrational spec-
troscopy. Chem. Phys. Lett. 408:284–89
149. Chen XK, Minofar B, Jungwirth P, Allen HC. 2010. Interfacial molecular organization at aqueous
solution surfaces of atmospherically relevant dimethyl sulfoxide and methanesulfonic acid using sum
frequency spectroscopy and molecular dynamics simulation. J. Phys. Chem. B 114:15546–53

216 Wang et al.


PC66-FrontMatter ARI 4 March 2015 12:22

Annual Review of
Physical Chemistry
Contents Volume 66, 2015

Molecules in Motion: Chemical Reaction and Allied Dynamics in


Solution and Elsewhere
James T. Hynes p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

Crystal Structure and Prediction


Tejender S. Thakur, Ritesh Dubey, and Gautam R. Desiraju p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p21
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

Reaction Dynamics in Astrochemistry: Low-Temperature Pathways to


Polycyclic Aromatic Hydrocarbons in the Interstellar Medium
Ralf I. Kaiser, Dorian S.N. Parker, and Alexander M. Mebel p p p p p p p p p p p p p p p p p p p p p p p p p p p p p43
Coherence in Energy Transfer and Photosynthesis
Aurélia Chenu and Gregory D. Scholes p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p69
Ultrafast Dynamics of Electrons in Ammonia
Peter Vöhringer p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p97
Dynamics of Bimolecular Reactions in Solution
Andrew J. Orr-Ewing p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 119
The Statistical Mechanics of Dynamic Pathways to Self-Assembly
Stephen Whitelam and Robert L. Jack p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 143
Reaction Dynamics at Liquid Interfaces
Ilan Benjamin p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 165
Quantitative Sum-Frequency Generation Vibrational Spectroscopy of
Molecular Surfaces and Interfaces: Lineshape, Polarization,
and Orientation
Hong-Fei Wang, Luis Velarde, Wei Gan, and Li Fu p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 189
Mechanisms of Virus Assembly
Jason D. Perlmutter and Michael F. Hagan p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 217
Cold and Controlled Molecular Beams: Production and Applications
Justin Jankunas and Andreas Osterwalder p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 241
Spintronics and Chirality: Spin Selectivity in Electron Transport
Through Chiral Molecules
Ron Naaman and David H. Waldeck p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 263

v
PC66-FrontMatter ARI 4 March 2015 12:22

DFT: A Theory Full of Holes?


Aurora Pribram-Jones, David A. Gross, and Kieron Burke p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 283
Theoretical Description of Structural and Electronic Properties of
Organic Photovoltaic Materials
Andriy Zhugayevych and Sergei Tretiak p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 305
Advanced Physical Chemistry of Carbon Nanotubes
Jun Li and Gaind P. Pandey p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 331
Site-Specific Infrared Probes of Proteins
Jianqiang Ma, Ileana M. Pazos, Wenkai Zhang, Robert M. Culik, and Feng Gai p p p p p 357
Biomolecular Damage Induced by Ionizing Radiation: The Direct and
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org

Indirect Effects of Low-Energy Electrons on DNA


Elahe Alizadeh, Thomas M. Orlando, and Léon Sanche p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 379
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

The Dynamics of Molecular Interactions and Chemical Reactions at


Metal Surfaces: Testing the Foundations of Theory
Kai Golibrzuch, Nils Bartels, Daniel J. Auerbach, and Alec M. Wodtke p p p p p p p p p p p p p p p p 399
Molecular Force Spectroscopy on Cells
Baoyu Liu, Wei Chen, and Cheng Zhu p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 427
Mass Spectrometry of Protein Complexes: From Origins
to Applications
Shahid Mehmood, Timothy M. Allison, and Carol V. Robinson p p p p p p p p p p p p p p p p p p p p p p p p p p 453
Low-Temperature Kinetics and Dynamics with Coulomb Crystals
Brianna R. Heazlewood and Timothy P. Softley p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 475
Early Events of DNA Photodamage
Wolfgang J. Schreier, Peter Gilch, and Wolfgang Zinth p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 497
Physical Chemistry of Nanomedicine: Understanding the Complex
Behaviors of Nanoparticles in Vivo
Lucas A. Lane, Ximei Qian, Andrew M. Smith, and Shuming Nie p p p p p p p p p p p p p p p p p p p p p 521
Time-Domain Ab Initio Modeling of Photoinduced Dynamics at
Nanoscale Interfaces
Linjun Wang, Run Long, and Oleg V. Prezhdo p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 549
Toward Design Rules of Directional Janus Colloidal Assembly
Jie Zhang, Erik Luijten, and Steve Granick p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 581
Charge Transfer–Mediated Singlet Fission
N. Monahan and X.-Y. Zhu p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 601
Upconversion of Rare Earth Nanomaterials
Ling-Dong Sun, Hao Dong, Pei-Zhi Zhang, and Chun-Hua Yan p p p p p p p p p p p p p p p p p p p p p p 619

vi Contents
PC66-FrontMatter ARI 4 March 2015 12:22

Computational Studies of Protein Aggregation: Methods and


Applications
Alex Morriss-Andrews and Joan-Emma Shea p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 643
Experimental Implementations of Two-Dimensional Fourier
Transform Electronic Spectroscopy
Franklin D. Fuller and Jennifer P. Ogilvie p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 667
Electron Transfer Mechanisms of DNA Repair by Photolyase
Dongping Zhong p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 691
Vibrational Energy Transport in Molecules Studied by
Relaxation-Assisted Two-Dimensional Infrared Spectroscopy
Natalia I. Rubtsova and Igor V. Rubtsov p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 717
Annu. Rev. Phys. Chem. 2015.66:189-216. Downloaded from www.annualreviews.org
Access provided by 143.255.218.155 on 05/21/20. For personal use only.

Indexes

Cumulative Index of Contributing Authors, Volumes 62–66 p p p p p p p p p p p p p p p p p p p p p p p p p p p 739


Cumulative Index of Article Titles, Volumes 62–66 p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 743

Errata

An online log of corrections to Annual Review of Physical Chemistry articles may be


found at http://www.annualreviews.org/errata/physchem

Contents vii

You might also like