Вы находитесь на странице: 1из 14

Methods 29 (2003) 196–209

www.elsevier.com/locate/ymeth

Vibrational Raman optical activity of proteins,


nucleic acids, and viruses
Ewan W. Blanch, Lutz Hecht, and Laurence D. Barron*
Department of Chemistry, University of Glasgow, Glasgow G12 8QQ, UK
Accepted 15 October 2002

Abstract

Due to its sensitivity to chirality, Raman optical activity (ROA), which may be measured as a small difference in vibrational
Raman scattering from chiral molecules in right- and left-circularly polarized incident light, is a powerful probe of biomolecular
structure in solution. Protein ROA spectra provide information on the secondary and tertiary structures of the polypeptide
backbone, hydration, side-chain conformation, and structural elements present in denatured states. Nucleic acid ROA spectra yield
information on the sugar ring conformation, the base stacking arrangement, and the mutual orientation of the sugar and base rings
around the C–N glycosidic linkage. ROA is able to simultaneously probe the structures of both the protein and the nucleic acid
components of intact viruses. This article gives a brief account of the theory and measurement of ROA and presents the ROA
spectra of a selection of proteins, nucleic acids, and viruses which illustrate the applications of ROA spectroscopy in biomolecular
research.
Ó 2003 Elsevier Science (USA). All rights reserved.

Keywords: Raman optical activity; Vibrational spectroscopy; Chirality; Protein conformation; Fold recognition; Conformational disease; Nucleic
acid structure; Virus structure

1. Introduction structure, dynamics, local environments and assembly


processes [1–8].
X-ray crystallography and high-resolution nuclear A particularly informative method of vibrational
magnetic resonance (NMR) spectroscopy are currently spectroscopy is vibrational Raman optical activity
preeminent in the field of structural biology due to their (ROA) which refers to a small difference in the intensity
ability to reveal the details of biomolecular structure at of Raman scattering from chiral molecules in right- and
atomic resolution and will continue to be invaluable. left-circularly polarized incident light or, equivalently, a
However, there are limitations to their applicability: small circularly polarized component in the scattered
many biomolecules have proven difficult to crystallize light [9,10]. This was first observed in 1973 [11] and has
while many others have structures too large to be cur- been developed in our laboratory to the point that it is
rently solved by NMR. Academic and commercial in- now an incisive probe of aqueous solution structures of
terests in the postgenomic era will demand structural peptides, proteins, carbohydrates, nucleic acids, and
information about such biomolecules. Although pro- intact viruses [12]. Conventional Raman spectroscopy
viding information at atomic resolution, vibrational provides molecular vibrational spectra by means of in-
spectroscopic techniques will become increasingly valu- elastic scattering of visible light. During the Stokes
able since they can be routinely applied to a wide range Raman scattering event, the interaction of a molecule
of biological systems and provide information on with an incident photon of energy gx (where x is its
angular frequency and g is DiracÕs constant) can leave
the molecule in an excited vibrational state of energy
gxm , with a corresponding loss of photon energy and
*
Corresponding author. Fax: +0141-330-4888. hence a shift to lower angular frequency x–xm of the
E-mail address: laurence@chem.gla.ac.uk (L.D. Barron). scattered photon. Therefore, by analyzing the scattered

1046-2023/03/$ - see front matter Ó 2003 Elsevier Science (USA). All rights reserved.
PII: S 1 0 4 6 - 2 0 2 3 ( 0 2 ) 0 0 3 1 0 - 9
E.W. Blanch et al. / Methods 29 (2003) 196–209 197

light with a visible spectrometer, a complete vibrational property tensors replaced by corresponding vibrational
spectrum may be obtained. In this article we provide a Raman transition tensors. For the case of a molecule
brief survey of the ROA technique, with some typical composed entirely of idealized axially symmetric bonds,
applications in biomolecular science illustrated by ex- for which bðG0 Þ2 ¼ bðAÞ2 and aG0 ¼ 0 [16,17], a simple
amples from recent studies on proteins, nucleic acids, bond polarizability theory shows that ROA is generated
and viruses. exclusively by anisotropic scattering and the CID ex-
pressions reduce to
Dð0°Þ ¼ 0
2. Basic theory
and
2
2.1. The ROA observables 32bðG0 Þ
Dð180°Þ ¼ h i:
2
c 45a2 þ 7bðaÞ
The fundamental scattering mechanism responsible
for ROA was discovered by Atkins and Barron [13]. A Unlike conventional Raman scattering intensities, which
more definitive version was developed by Barron and are the same in forward and backward directions, ROA
Buckingham [14], who introduced the following defini- intensity is therefore maximized in backscattering and
tion of the dimensionless circular intensity difference zero in forward scattering.
(CID),
D ¼ ðI R  I L Þ=ðI R þ I L Þ; ð1Þ 2.2. Enhanced sensitivity of ROA to structure and
as an appropriate experimental observable, where I and R dynamics of chiral biomolecules
I L are the scattered intensities in right- and left-circularly
polarized incident light, respectively. With regard to the The normal vibrational modes of biopolymers can be
electric dipole–electric dipole molecular polarizability highly complex as they contain contributions from local
tensor aab and the electric dipole–magnetic dipole and vibrational coordinates in both the backbone and the
the electric dipole–electric quadrupole optical activity side chains. ROA is able to cut through the complexity
tensors G0ab and Aabc [15,16], the CIDs for forward (0°) of the corresponding vibrational spectra as the largest
and backward (180°) scattering from an isotropic sam- ROA signals originate from vibrational coordinates
ple for incident wavelengths much greater than the which sample the most rigid and chiral parts of the
molecular dimensions are structure. These usually reside within the backbone and
h i give rise to ROA band patterns characteristic of the
2 2
4 45aG0 þ bðG0 Þ  bðAÞ backbone conformation. In proteins these include bands
Dð0°Þ ¼ h i ð2Þ characteristic of secondary, loop, and turn structures. In
2
c 45a2 þ 7bðaÞ comparison, the parent conventional Raman spectrum
and of a protein is dominated by bands arising from the
h  2 i amino acid side chains which may obscure the peptide
2
24 b G0 þ 13 bðAÞ backbone bands. A few distinct ROA signals from the
Dð180°Þ ¼ h i ð3Þ amino acid side chains, although less prominent than
2
c 45a2 þ 7bðaÞ those from the peptide backbone, can also provide
useful information. Carbohydrate ROA spectra are
where the isotropic invariants are defined as
similarly dominated by signals from skeletal vibrations,
1 1 principally the constituent sugar rings and the connect-
a ¼ aaa and G0 ¼ G0aa
3 3 ing glycosidic links. The ROA spectra of nucleic acids
and the anisotropic invariants as are dominated by bands characteristic of the stereo-
1  chemical arrangement of the bases with respect to each
2
bðaÞ ¼ 3aab aab  aaa abb ; other, to sugar rings, and to signals from the sugar–
2
phosphate backbone.
1  The time scale of the Raman scattering event
bðG0 Þ2 ¼ 3aab G0ab  aaa G0bb ; (3:3  1014 s for a vibration with Stokes wavenumber
2
and shift 1000 cm1 excited in the visible) is much shorter
than that of the fastest conformational fluctuations.
1 Therefore the ROA spectrum is a superposition of
bðAÞ2 ¼ xaab eacd Acdb :
2 ‘‘snapshot’’ spectra from all the distinct conformations
These results apply specifically to Rayleigh (elastic) present in the sample at equilibrium. ROA intensity is
scattering. For Raman (inelastic) scattering the same dependent on absolute chirality and therefore yields a
basic CID expressions apply but with the molecular cancellation of contributions from enantiomeric struc-
198 E.W. Blanch et al. / Methods 29 (2003) 196–209

tures which can arise as a mobile structure explores the signals, the spectral acquisition is synchronized with an
range of accessible conformations. These factors result electro-optic modulator used to switch the state of po-
in ROA exhibiting an enhanced sensitivity to the dy- larization of the incident laser beam between right- and
namic behavior of biomolecular structure. In contrast, left-circular at a suitable rate.
observables that are ‘‘blind’’ to chirality, such as con- Spectra are displayed in analog-to-digital converter
ventional Raman scattering intensities, are generally units as a function of the Stokes Raman wavenumber
additive and therefore less sensitive to conformational shift with respect to the exciting laser line. The green
mobility. Ultraviolet circular dichroism (UVCD) also 514.5-nm line of an argon ion laser is used as this pro-
shows an enhanced sensitivity to the dynamics of chiral vides a compromise between reduced fluorescence from
structures. However, the sensitivity is less than that for sample impurities with increasing wavelength and in-
ROA due to its dependence on electronic transitions. creased scattering efficiency with decreasing wavelength
due to the Rayleigh k4 law. Typical laser power at the
sample is 700 mW, sample concentrations of proteins,
3. Instrumentation polypeptides, and nucleic acids are 30–100 mg/mL,
and those of intact viruses are 5–30 mg/mL. Under
Since ROA is maximized in backscattering, a back- these conditions ROA spectra such as those presented
scattering geometry has proven to be essential for the here may be obtained in 5–24 h for proteins and nu-
routine measurement of ROA spectra of biomolecules in cleic acids and 1–4 days for intact viruses. Measure-
aqueous solution. The optical layout of our current ments can also be performed over the temperature range
backscattering ROA instrument is shown in Fig. 1 [18]. of 0–60 °C by directing dry air downward over the
A visible argon ion laser beam is weakly focused into the sample cell from a device used to cool protein crystals in
sample solution which is contained in a small rectan- X-ray diffraction experiments, to study dynamic be-
gular fused quartz cell. The cone of backscattered light havior.
is reflected off a 45 ° mirror, which has a small central Several novel design features of a new ROA instru-
hole drilled to allow passage of the incident laser beam, ment developed by Hug and Hangartner [19] will be
through an edge filter to remove the Rayleigh line and especially valuable for biomolecular studies. It is based
into the collection optics of a single grating spectro- on measuring the circularly polarized component in the
graph. This is a customized version of a fast-imaging Raman scattered light instead of the Raman intensity
spectrograph containing a highly efficient volume holo- difference in right- and left-circularly polarized incident
graphic transmission grating. The detector is a cooled light employed in the Glasgow instrument. This allows
back-thinned charge-coupled device (CCD) camera with ‘‘flicker noise’’ arising from dust particles, density fluc-
a quantum efficiency of 80% over the spectral range of tuations, etc., to be eliminated because the intensity
our studies. Operation of the CCD camera in multi- differences required to extract the circularly polarized
channel mode allows the full spectral range to be mea- components of the scattered beam are taken between
sured in a single acquisition. To measure the small ROA two components of the scattered light measured during

Fig. 1. Optical layout of the current backscattering ROA instrument used in Glasgow.
E.W. Blanch et al. / Methods 29 (2003) 196–209 199

the same acquisition period. The flicker noise therefore


cancels out, resulting in superior signal-to-noise char-
acteristics. A commercial instrument based on this de-
sign will be available shortly.

4. Studying biomolecules

4.1. Proteins

4.1.1. Folded proteins


Because protein ROA spectra are usually dominated
by bands originating in the peptide backbone they di-
rectly reflect the solution conformation. The enhanced
sensitivity of ROA to dynamic aspects of structure also
makes it a useful new probe of order–disorder transi-
tions and conformational mobility.
Vibrations of the peptide backbone in proteins are
usually associated with three main regions of the Raman
spectrum [2,3,20]: the backbone skeletal stretch region
870–1150 cm1 which arises from Ca AC; Ca ACb and
Ca AN stretch coordinates; the extended amide III re-
gion 1230–1340 cm1 which involves mainly the in-
phase combination of the N–H in-plane deformation
with the Ca AN stretch along with mixing between the
N–H and the Ca AH deformations; and the amide I re-
gion 1630–1700 cm1 which originates in the C@O
stretch. The extended amide III region is particularly
important for ROA studies as the coupling between the
NAH and Ca AH deformations is sensitive to geometry
and generates a rich and informative band structure [21].
Fig. 2a shows the backscattered Raman and ROA
spectra of the a-helical protein human serum albumin.
The main features of the ROA spectrum are in good
agreement with the Protein Data Bank (PDB) X-ray
crystal structure 1ao6, which reports 69.2% a-helix and
1.7% 310 -helix with the remainder being made up of
loops and turns. The strong sharp positive band at
1340 cm1 is assigned to a hydrated form of a-helix
while the weaker positive band at 1300 cm1 appears
to be associated with a-helix in a more hydrophobic
environment. We have reported similar bands arising Fig. 2. Backscattered Raman (I R þ I L ) and ROA (I R  I L ) spectra of
(a) human serum albumin, (b) jack bean concanavalin A, and (c) hen
from elements of a-helix in the ROA spectra of model a- lysozyme at pH 5.4 and 20 °C.
helical polypeptides and globular proteins [12,22,23] and
the coat proteins of intact filamentous bacteriophages
[24]. The relative intensities of these two bands appear to wavenumbers lower). The positive ROA band from a-
correlate with the exposure of the polypeptide backbone helix at 1300 cm1 retains much of its intensity in D2 O.
to the solvent within the elements of a-helix in each case. Conventional Raman bands at similar wavenumbers
For example, the positive a-helix protein ROA band at have been assigned to a-helix in studies performed on
1340 cm1 completely disappears when the protein is polypeptides [25] and filamentous bacteriophages [26].
dissolved in D2 O [12]. This indicates both that the cor- In particular, a recent Raman study of a-helical poly(L -
responding sequences are exposed to solvent and that alanine) [25] has provided definitive assignments of a
N–H deformations of the peptide backbone make a number of the normal modes of vibration in the ex-
significant contribution to the generation of this band tended amide III region over the range in which we have
(because corresponding N–D deformations contribute identified ROA bands. These normal modes variously
to normal modes in a spectral region several hundred transform the same as the A, E1 , and E2 symmetry
200 E.W. Blanch et al. / Methods 29 (2003) 196–209

species of the point group of a model infinite regular may also contribute in this region: in fact the positive
helix. The ROA bands assigned to a-helix in this region component at 1295 cm1 of this b-sheet couplet in
may be related to a number of these normal modes, with concanavalin A may originate in vibrations of the b-
the ROA intensities and exact wavenumbers being a turns in the hairpin bends present in the up-and-down
function of the perturbations (geometric and/or due to multistranded b-sheet of this protein. Amide III bands
various types of hydration) to which the particular he- from b-sheet in conventional Raman spectroscopy are
lical sequence is subjected. assigned to the region 1230– 1245 cm1 [3]. The amide
Insight into the nature of the hydrophobic and hy- I couplet, negative at 1658 cm1 and positive at 1677
drated variants of a-helix has been provided by recent cm1 , is another signature of b-sheet and can be easily
electron spin resonance studies of double spin-labeled distinguished from the amide I couplet produced by a-
alanine-rich peptides which identified a new, more open helix, which typically occurs 5–20 cm1 lower. This
conformation of the a-helix [27,28]. Computer model- correlates with b-sheet amide I bands in conventional
ing indicates that this more open geometry leaves the Raman spectroscopy which occur in the range
hydrogen bonding intact but changes the C@O N 1665–1680 cm1 [3]. A number of ROA bands associ-
angle, resulting in the splaying of the backbone amide ated with b-sheet may appear in the backbone skeletal
carbonyls away from the helix axis and into the solu- stretch region, as found here for concanavalin A.
tion. Therefore, the more open structure may be the However, the details of wavenumber, intensity, and
preferred conformation in aqueous solution as it allows band shape are variable, possibly reflecting differences in
hydrogen bonding with water molecules. An equilib- local conformations and amino acid compositions found
rium would then exist between the canonical form of a- within different b-sheets.
helix, which might be responsible for the positive ROA Proteins containing a significant amount of b-sheet
band at 1300 cm1 , and the more open form, which often display additional ROA signals originating in
may generate that at 1340 cm1 , with the former be- loops and turns. Negative ROA bands in the range
ing favored in a hydrophobic environment and the 1340–1380 cm1 appear to originate in b-hairpin
latter in a hydrophilic (or hydrogen-bonding) environ- bends; an example is a band of medium intensity ap-
ment. pearing at 1345 cm1 in the spectrum of concanavalin
The amide I ROA couplet, negative at 1640 cm1 A. This signature allows ROA to differentiate between
and positive at 1665 cm1 , is also a characteristic sig- parallel and antiparallel types of b-sheet as only the
nature of a-helix and corresponds to the wavenumber latter usually contain hairpin bends while the ends of
range 1645–1655 cm1 for a-helix bands in conven- strands in the former are usually connected by a-helical
tional Raman spectra [3]. Positive ROA intensity in the sequences. Many b-sheet proteins also show a strong
range 870–950 cm1 is a further signature of a-helix positive ROA band at 1314–1325 cm1 similar to the
with the detailed band structure in this region appearing prominent bands observed in disordered forms of
to show a dependence on side-chain composition, helix poly(L -lysine) and poly(L -glutamic acid) [12,29]. These
length, and the presence of irregularities. disordered polypeptides are thought to contain signifi-
Fig. 2b shows the Raman and ROA spectra of the b- cant amounts of the polyproline II (PPII) helix confor-
sheet protein jack bean concanavalin A which contains, mation. This signal may therefore be a signature of the
according to the PDB X-ray crystal structure 2cna, PPII helical elements known from X-ray crystal struc-
43.5% b-strand, 1.7% a-helix, and 1.3% 310 -helix in a tures to occur in some of the longer loops between ele-
jelly roll b-barrel with the remainder being hairpin bends ments of secondary structure [30,31]. An example of
and long loops. The sharp negative band at 1241 cm1 such a PPII signal is observed at 1316 cm1 in the
has been assigned to b-structure. Similar bands have ROA spectrum of jack bean concanavalin A.
been observed in the ROA spectra of other proteins Fig. 2c shows the Raman and ROA spectra of the
containing b-sheet [12,23], which sometimes has another a þ b protein hen lysozyme. The fold of this protein is
negative band at 1220 cm1 which appears to be as- very different from those of human serum albumin and
sociated with a distinct variant of b-strand or sheet, concanavalin A mentioned above. This is reflected in the
possibly hydrated. As more ROA data have been accu- large differences between their ROA spectra. Hen lyso-
mulated on proteins and polypeptides containing b-sheet zyme contains 28.7% a-helix, 10.9% 310 -helix, and 6.2%
it has become apparent that the true signature of some b-sheet according to the PDB X-ray crystal structure
types of b-sheet in the extended amide III region may be 1lse, which is consistent with the presence of the positive
a couplet, negative at low wavenumber and positive at ROA bands assigned to hydrophobic and hydrated a-
high. The negative peak of the couplet in proteins ap- helix at 1299 and 1342 cm1 , respectively, and with the
pears to be constrained either to the region 1220 cm1 sharp negative band at 1240 cm1 indicative of b-
or the region 1240 cm1 . The positive peak of this structure. As our database of protein ROA spectra has
couplet appears to be more variable and appears in the expanded it has become apparent that the large positive
range 1260–1295 cm1 . Bands from loops and turns peak 1297–1300 cm1 observed for proteins with a
E.W. Blanch et al. / Methods 29 (2003) 196–209 201

lysozyme-type fold may be boosted by other bands. is possible to deduce motif types of supersecondary
Proteins with a lysozyme-type fold often contain an elements and even the domain fold class from ROA
unusually high 310 -helix component. It is possible that spectra. We are developing a pattern recognition pro-
the positive band 1299 cm1 contains a contribution gram, based on principal component analysis (PCA), to
from a 310 -helix band, with a positive signal identify protein folds from ROA spectral band patterns
1295 cm1 . There may also be bands contributing in [34]. From the ROA data, the PCA program calculates a
this region from turns. A definitive analysis of all ROA set of subspectra that serve as basis functions. The al-
bands in the extended amide III region has not been gebraic combination of these subspectra, weighted with
possible to date due to the difficulty of deconvoluting appropriate expansion coefficients, can be used to re-
the complex, bisignate spectra. construct the original set of experimental ROA spectra.
The amide I couplet, negative at 1641 cm1 and Our current set contains over 70 entries comprising
positive at 1665 cm1 with a small shoulder at ROA spectra of several polypeptides in model confor-
1683 cm1 , also indicates the presence of a-helix and a mations, many proteins with well-defined folds known
lesser amount of b-sheet. There is a small couplet, neg- from X-ray crystallographic or solution NMR mea-
ative at 1426 cm1 and positive at 1462 cm1 , from surements, and several viruses with coat protein folds
CH2 and CH3 side-chain deformations. On the low- known from X-ray crystallographic or fiber diffraction
wavenumber side of this couplet there is another rela- data.
tively weak couplet, positive at low wavenumber and
negative at high wavenumber, that originates in tryp- 4.1.2. Unfolded proteins
tophan vibrations. In addition, there is a relatively large As yet it has not been possible to obtain useful ROA
positive band 1554 cm1 assigned to the W3-type vi- spectra of fully unfolded denatured states of proteins
brational mode of the indole ring of tryptophan resi- which have well-defined tertiary folds in the native state.
dues. Miura et al. [32] found that the magnitude of the This is because the intense Raman bands from chemical
torsion angle v2;1 of the tryptophan side chain, which denaturants typically used at high concentration pre-
describes the orientation of the indole ring with respect clude ROA measurements, while thermally unfolded
to the local peptide backbone, can be deduced from the proteins often give rise to intense Rayleigh light scat-
wavenumber of the corresponding conventional Raman tering due to aggregation. We have, however, reported
band. Similarly, the magnitude of the torsion angle can interesting results for partially unfolded denatured
be determined from the position of the W3-type vibra- protein states associated with molten globules and re-
tional mode in the ROA spectrum but with the added duced proteins and for proteins which are unfolded in
advantage of a possible determination of its sign from their native biologically active states [35].
the measured sign of the ROA band and hence the de- Molten globules are partially unfolded denatured
duction of the absolute stereochemistry of the trypto- protein states, stable at equilibrium, with well-defined
phan side chain [33]. This information is not normally secondary structure but lacking the specific tertiary in-
available other than from a structure determined at teractions characteristic of the native state [36,37]. A
atomic resolution. Hen lysozyme contains six trypto- much studied molten globule is that supported by a-
phan residues, four with positive v2;1 values and two lactalbumins at low pH and called the A-state. Native
with negative v2;1 values according to the PDB structure bovine a-lactalbumin has the same fold and X-ray crystal
1lse. Partial cancellation of the resulting ROA signals structure very similar to that of hen lysozyme and this is
yields the positive band observed in Fig. 2c [33]. In this apparent from its ROA spectrum [22] shown in Fig. 3a,
manner the W3 ROA band can also be used as a probe which is similar to that of hen lysozyme depicted in Fig.
of conformational heterogeneity among tryptophan 2c. Nevertheless, differences in detail the spectra are
residues in disordered protein sequences as the cancel- apparent. This highlights the sensitivity of ROA to small
lation of signals with opposing signs results in a loss of differences in structure in proteins with the same fold.
ROA intensity. This is similar to the disappearance of The Raman and ROA spectra of A-state bovine a-lact-
near-UVCD bands from aromatic residues upon the albumin are shown in Fig. 3b. The sensitivity of ROA to
loss of tertiary structure. Thus these two techniques the complexity of order in molten globule states is indi-
provide complementary views of order/disorder transi- cated by the large differences observed here between the
tions as ROA probes the intrinsic skeletal chirality of ROA spectra of the native and A-states of bovine a-
the tryptophan side chain while UVCD probes the lactalbumin as opposed to the small differences between
chirality in the immediate vicinity of the aromatic the corresponding parent Raman spectra. Much of the
chromophore. ROA band structure in the extended amide III region of
Since ROA spectra contain bands characteristic of the A-state spectrum has disappeared and is replaced by
loops and turns and bands indicative of secondary a large couplet consisting of a single negative band ob-
structure, they provide information about the overall served at 1236 cm1 and two distinct positive bands at
three-dimensional solution structure of a protein. It 1297 and 1312 cm1 . The negative 1236 cm1 signal
202 E.W. Blanch et al. / Methods 29 (2003) 196–209

nate in reduced hydration in some a-helical sequences. In


the amide I region the positive signal of the couplet has
shifted by 10 cm1 to higher wavenumber compared to
that for the native state. The signal originating from the
W3 vibrational mode of tryptophan residues at
1551 cm1 has almost completely disappeared in the
ROA spectrum of the A-state and been replaced by a
weak positive band at 1560 cm1 indicating confor-
mational heterogeneity among these side chains associ-
ated with a loss of the characteristic tertiary interactions
found in the native state.
The Raman and ROA spectra of reduced hen lyso-
zyme are shown in Fig. 3c [35]. This sample was prepared
by reducing all the disulfide bonds and keeping the
sample at low pH to prevent their reoxidation. There are
significant changes of the ROA spectrum of reduced ly-
sozyme compared with that of the native state shown in
Fig. 2c. The bands in the backbone skeletal stretch region
are suppressed, along with differences in detail, indicat-
ing the loss of much of the fixed structure. Most of the
ROA band patterns in the extended amide III region
have disappeared and have been replaced by a large
broad couplet with some hints of weak band structure,
similar to that observed in the ROA spectrum of A-state
a-lactalbumin, suggesting that the reduced protein sup-
ports a number of conformations with a range of local
residue /; w angles clustering in the same regions of the
Ramachandran surface as those in the native state. The
disappearance of the positive band at 1340 cm1 im-
plies that none of the significant amount of hydrated a-
helix found in the native state persists in the reduced
form. The positive signal of the amide I couplet has also
shifted by 9 cm1 to higher wavenumber and has be-
come quite sharp, indicating the presence of a significant
amount of b-structure. The couplet from 1426–
1462 cm1 in the spectrum of the native state originating
in aliphatic side chains and the band at 1554 cm1 as-
signed to tryptophan side chains are all greatly dimin-
ished, presumably due to conformational heterogeneity.
Although of similar appearance, the ROA spectra of
Fig. 3. Backscattered Raman and ROA spectra of (a) native bovine a-
different denatured proteins display differences in detail,
lactalbumin at pH 4.6 and 20 °C, (b) A-state bovine a-lactalbumin at
pH 1.9 and 2 °C, and (c) reduced hen lysozyme at pH 2.4 and 20 °C. possibly reflecting the different residue compositions and
their different /; w propensities.
Although human lysozyme has a stability similar to
may originate in residues clustering around an ‘‘average’’ that of hen lysozyme with regard to temperature and
of the conformations corresponding to the two negative low pH, its thermal denaturation behavior is subtly
signals in the native spectrum, observed at 1222 and different. Below pH 3.0 the thermal denaturation of
1246 cm1 in Fig. 3a, assigned to b-structure. The posi- human lysozyme is not a two-state process. Unlike hen
tive band at 1297 cm1 may arise from the same a-he- lysozyme, it supports a partially folded molten globule-
lical sequences in a hydrophobic environment, and like state at elevated temperatures [38]. Incubation at
possibly 310 -helical sequences, as those responsible for a 57 °C and pH 2.0, under which conditions the partially
similar band in the spectrum of the native state. The folded state is the most highly populated, induces the
positive 1340 cm1 band assigned to a-helix in a hy- formation of amyloid fibrils while incubation at 70 °C
drated state that dominates the ROA spectrum of the and pH 2.0, under which conditions the fully denatured
native state is completely lost in the A-state, with a new state is the most highly populated, leads to the forma-
positive band appearing at 1312 cm1 that may origi- tion of amorphous aggregates [39].
E.W. Blanch et al. / Methods 29 (2003) 196–209 203

The Raman and ROA spectra of the native and has undergone a conformational change to PPII struc-
partially folded prefibrillar–amyloidogenic intermediate ture. The disappearance of the positive 1550 cm1
states of human lysozyme are shown in Fig. 4a and b, band assigned to tryptophan vibrations indicates that
respectively [40]. The structure and ROA spectrum of major conformational changes have occurred among the
the native state of human lysozyme are similar to those five tryptophan residues, four of which lie within the a-
of hen lysozyme. However, large changes are apparent helical domain. Thus the ROA data suggest that the a-
in the ROA spectrum of the prefibrillar intermediate. domain destabilizes and undergoes a conformational
The most significant of these is the loss of the positive change in the prefibrillar intermediate and that PPII
1345 cm1 band assigned to hydrated a-helix and the helix may be a critical conformational element involved
appearance of a new positive band at 1325 cm1 as- in amyloid fibril formation [40]. There is no sign of an
signed to PPII helix. This suggests that hydrated a-helix increase in b-sheet content in the intermediate. The
ROA spectrum of hen lysozyme, which has a much
lower propensity to form amyloid fibrils than the human
variant [41], is virtually native-like under the same
conditions of low pH and elevated temperature [40].
ROA has proven useful in several recent studies of
natively unfolded proteins of biological and physiolog-
ical significance. The ROA spectra of bovine milk
caseins [42], several wheat prolamins [35], and the hu-
man recombinant synuclein and tau brain proteins,
some of which are associated with neurodegenerative
disease [42], were all found to be dominated by a strong
positive band at 1316–1322 cm1 assigned to PPII
helix. The ROA spectrum of c -synuclein shown in Fig.
4c displays an example at 1321 cm1 . The absence of a
well-defined amide I ROA couplet in this spectrum in-
dicates the lack of secondary structure. Although the
other natively unfolded proteins studied so far have
ROA spectra that are similar overall, there are differ-
ences of detail which reflect differences in residue com-
position and minor differences in structural elements.
The studies on the caseins, synucleins, and tau suggested
that these proteins may be more realistically classified as
being ‘‘rheomorphic,’’ meaning flowing shape [43], ra-
ther than ‘‘random coil.’’ The rheomorphic state is dis-
tinct from the molten globule state which is a more
compact entity containing a hydrophobic core and a
significant amount of secondary structure [36,37].
The conformational plasticity supported by mobile
regions within native proteins, partially denatured pro-
tein states, and natively unfolded proteins underlies
many of the conformational (protein misfolding) dis-
eases [44,45], many of which involve amyloid fibril for-
mation. As it is extended, flexible, and hydrated, the
PPII helical structure identified in some of these systems
imparts a plastic open character to the structure and
may be implicated in the formation of regular fibrils in
the amyloid diseases [40,42]. This is because the elimi-
nation of water molecules between extended polypeptide
chains with fully hydrated N–H and C@O groups to
form b-sheet hydrogen bonds is a highly favorable
process entropically [46]. As the PPII- and b-sheet re-
gions are adjacent on the Ramachandran surface it is
Fig. 4. Backscattered Raman and ROA spectra of native human ly-
sozyme at pH 5.4 and 20 °C, (b) the prefibrillar intermediate of hu- expected that elements of PPII helix would readily un-
man lysozyme at pH 2.0 and 57 °C, and (c) human c-synuclein at dergo this type of aggregation with each other and with
pH 7.0 and 20 °C. the edges of preexisting b-sheet to form the cross-b
204 E.W. Blanch et al. / Methods 29 (2003) 196–209

structures typical of amyloid fibrils. True random coil is


expected to lead to amorphous aggregation rather than
fibril formation and this is what is generally observed.
However, although the presence of significant amounts
of PPII structure may be necessary for the formation of
regular fibrils, other factors might be important too
because, of the milk and brain proteins studied by ROA,
only a-synuclein and tau are fibrillogenic and associated
with disease. It appears necessary to consider properties
of the constituent residues such as mean hydrophobicity
and charge also [42].
PPII helix is not readily amenable to traditional
methods of structure determination due to its inherent
flexibility and the lack of intrachain hydrogen bonding,
which has hindered its recognition in globular proteins.
For instance, the PPII conformation is indistinguishable
from an irregular backbone structure for free peptides in
solution by 1 H NMR spectroscopy [47]. PPII helix may
be recognized in polypeptides by UVCD [47–49] and
vibrational circular dichroism [7,50,51] and in proteins
from the deconvolution of UVCD spectra [52]. In con-
trast, ROA provides a clear and characteristic signature
of PPII helix in both proteins and polypeptides.

4.2. Nucleic acids

The study of the structure and function of nucleic


acids remains a central problem in molecular biology.
Although ROA studies of nucleic acids are at an early
stage, the results obtained so far are promising as ROA
appears to be sensitive to three different sources of chi-
rality: the chiral base-stacking arrangement of intrinsi-
cally achiral base rings, the chiral disposition of the base
and sugar rings with respect to the C–N glycosidic link,
and the inherent chirality associated with the asym-
metric centres of the sugar rings. Model studies on
pyrimidine nucleosides [53] and polyribonucleotides [54–
56] have provided a basis for the interpretation of ROA
spectra of DNA and RNA.
The Raman and ROA spectra of calf thymus DNA
and of phenylalanine-specific transfer RNA (tRNAPhe ) Fig. 5. Backscattered Raman and ROA spectra of (a) calf thymus
in the presence and absence of Mg2þ ions are shown in DNA, (b) Mg2þ -bound tRNAPhe , and (c) Mg2þ -free tRNAPhe at pH
Fig. 5a–c, respectively [57]. ROA bands in the region 6.8 and 20 °C.
900–1150 cm1 originate in vibrations of the sugar
rings and phosphate backbone. The region 1200– puckers are mainly C20 -endo and the RNAs taking up
1550 cm1 is dominated by normal modes in which the A-type double helical segments where the sugar puckers
vibrational coordinates of the base and sugar rings are are mainly C30 -endo. There are smaller differences be-
mixed. ROA band patterns in this sugar–base region tween the two RNA spectra and these are most apparent
appear to reflect the mutual orientation of the two rings in the sugar–phosphate region 900– 1150 cm1 . It is
and possibly the sugar ring conformation. The region known that Mg2þ ions are necessary to hold RNAs in
1550–1750 cm1 contains ROA bands characteristic of their specific tertiary folds. For tRNAPhe in the presence
the types of bases involved and the particular stacking of Mg2þ this is a compact L-shaped form, whereas in the
arrangements. Although the ROA spectra of the DNA absence of Mg2þ the tRNAPhe adopts an open cloverleaf
and the two RNAs are similar there are many differences structure [58]. The ROA spectrum of the Mg2þ -free
in detail. The main differences originate in the DNA tRNAPhe shows a strong negative, positive, negative
taking up a B-type double helix in which the sugar triplet at 992, 1048, and 1091 cm1 which is very
E.W. Blanch et al. / Methods 29 (2003) 196–209 205

similar to that found in A-type polyribonucleotides and


has been assigned to the C30 -endo sugar pucker. This
signature is weaker and more complex in the ROA
spectrum of the Mg2þ -bound sample, suggesting a wider
range of sugar puckers. Switching the sugar pucker
conformations from C30 -endo to C20 -endo would elon-
gate the sugar–phosphate backbone and may assist the
formation of the loops and turns that characterize the
tertiary structure of the folded form.

4.3. Viruses

Knowledge of the structure of viruses at the molecular


level is essential for enterprises such as structure-guided
antiviral drug design [59]. However, the application of
key structural biology techniques such as X-ray crystal-
lography or fiber diffraction is often hampered by prac-
tical difficulties. Conventional Raman is valuable in
studies of intact viruses at the molecular level as it is able
to simultaneously probe both the protein and the nucleic
acid constituents [3,6]. The additional incisiveness of
ROA further enhances the value of Raman spectroscopy
in structural virology.
The first virus ROA spectra were reported for fila-
mentous bacteriophages [24]. The Raman and ROA
spectra of Pf1 [24], fd, and PH75 [33] are shown in Figs.
6a–c, respectively, as examples. At high concentrations
(above 10–20 mg/mL depending on virus and detailed
experimental conditions) filamentous bacteriophages
form liquid crystals which give rise to artifacts which
prevent the acquisition of reliable ROA data. Hence the
viral concentrations were kept below the liquid crystal-
forming thresholds which consequently resulted in a
decrease of signal-to-noise levels. This was partly offset
by an apparent boost in viral protein ROA CID values,
possibly due to decreased conformational mobility
within closely packed environments. The water back-
ground has been subtracted from each of the corre-
sponding Raman spectra shown in Fig. 6 to aid
presentation.
Filamentous bacteriophages are long flexible rods Fig. 6. Backscattered Raman and ROA spectra of the filamentous
made up of a loop of single-stranded DNA surrounded bacteriophages (a) Pf1, (b) fd, and (c) pH 7.5 at pH 7.6 and 20 °C.
by a shell consisting of a helical array of several thou-
sand identical copies of a major coat protein containing tive intensities of the ROA bands at  1300 and
50 amino acids. There are also a few copies of minor 1342 cm1 assigned to hydrophobic and hydrated a-he-
coat proteins. X-ray fiber diffraction has shown that the lix, respectively, correlate well with the hydration of the
major coat protein subunits have the fold of an extended peptide backbone in each coat protein [24]. The interior,
a-helix and that they overlap each other so that the protected region of the coat protein of Pf1 contains a
exterior half of each protein is exposed to water while continuous sequence of 20 hydrophobic residues while
the interior half is protected by neighboring coat pro- the exterior, exposed region contains a 20-residue se-
teins [60]. The ROA spectra are dominated by bands quence of mixed hydrophobic and hydrophilic residues.
assigned to a-helix in the backbone skeletal stretch, In Fig. 6a the maximum intensities of the corresponding
amide I, and extended amide III regions which have ROA bands at 1299 and 1343 cm1 are virtually
been discussed already. A couplet, negative at 1095
identical. For bacteriophage fd the sequence of pure
5 cm1 and positive at 1125
5 cm1 , also seems to be hydrophobic helix in the interior region of the coat
associated with well-defined a-helix. Further, the rela- protein is shorter and there is a longer sequence of
206 E.W. Blanch et al. / Methods 29 (2003) 196–209

mixed hydrophobic and hydrophilic residues. In the


corresponding ROA spectrum in Fig. 6b the hydrated a-
helix band is now relatively stronger than the band
originating in hydrophobic a-helix.
An obvious difference between the three bacterio-
phage spectra is the behavior of the band associated
with the W3-type vibrational mode of tryptophan resi-
dues. The major coat protein of Pf1 does not contain
any tryptophan residues and there is no ROA signal
1550 cm1 . Bacteriophage fd contains a single trypto-
phan residue in the major coat protein which generates
the positive ROA band at 1559 cm1 while PH75,
which has unusual thermophilic properties [61], shows a
negative band at 1550 cm1 originating from the single
tryptophan in its coat protein. This suggests that the
absolute stereochemistry of the indole ring of the tryp-
tophan residue relative to the local peptide backbone in
fd is quasienantiomeric to that in PH75. From the
wavenumbers and signs of the W3 ROA bands it was
deduced that v2;1 ¼ þ120 ° for fd and )93° for PH75
[33].
Fig. 7a–c show, respectively, the Raman and ROA
spectra of tobacco mosaic virus (TMV) [62], a tob-
amovirus, and the potexviruses potato virus X (PVX)
and narcissus mosaic virus (NMV) [34]. These are helical
plant viruses containing a single strand of positive-sense
RNA encapsidated within a rigid rod-shaped particle in
the case of TMV or a flexuous filamentous particle in the
case of the potexviruses, made up of multiple copies of a
single-coat protein. The structure of the coat protein
subunits of TMV has been determined by X-ray fiber
diffraction [63,64] to be based on a four-helix bundle
motif which contains both water-exposed residues and
residues at the hydrophobic interfaces between a-helices.
The ROA band pattern in the extended amide III region
of the TMV spectrum is characteristic of a helix bundle,
with a positive band 1295 cm1 assigned to a-helix in a
hydrophobic environment and a slightly stronger band
1342 cm1 assigned to hydrated a-helix. The amide I
couplet and bands in the backbone skeletal stretch re-
gion are also characteristic of a-helix. Fig. 7. Backscattered Raman and ROA spectra of the helical plant
Little is known about the conformations of the coat viruses (a) TMV, (b) PVX, and (c) NMV at pH 7.4 and 20 °C.
protein subunits of PVX and NMV from other tech-
niques but the similarities of their ROA spectra to the
ROA spectrum of TMV reveal they are both based on icosahedral virus with a single strand of RNA inside a
helix bundle folds. All three ROA spectra also contain a capsid composed of 60 identical copies of a coat protein
positive band at 1316 cm1 assigned to elements of with 159 amino acids. The X-ray crystal structure of
PPII helix in long loops and a small negative band at STMV [65] reveals the fold of the coat protein to be a
1220 cm1 from segments of b-strand, possibly hy- jelly roll b-barrel. Although it is not shown here, the
drated. Although the three ROA spectra are similar, the ROA spectrum of STMV is very similar to that of
differences in detail illustrate the sensitivity of ROA to proteins with a jelly roll fold such as concanavalin A,
different structural parameters and types of subunit shown in Fig. 2b. MS2 is a small T ¼ 3 icosahedral virus
packing within different virus capsids. with a single strand of RNA contained inside a capsid
In addition to filamentous and rod-shaped viruses, made up of 180 identical copies of a coat protein con-
ROA has also been applied to icosahedral viruses. Sa- taining 129 amino acids. The X-ray crystal structure of
tellite tobacco mosaic virus (STMV) is a small T ¼ 1 MS2 [66] reveals that the coat protein subunits fold into
E.W. Blanch et al. / Methods 29 (2003) 196–209 207

a 5-stranded antiparallel b-sheet with an additional


short hairpin at the N terminus and two a-helices. The
a-helices are responsible for interactions with a second
subunit to form a dimer containing 10 adjacent anti-
parallel b-strands. This fold is quite distinct from those
of the jelly roll type of STMV and most other icosahe-
dral viruses, and the ROA spectrum of the MS2 capsid
was found to reflect this.
Determination of the structures of the nucleic acid
component of intact viruses has proven difficult in even
the best-resolved X-ray crystal structures. Several of the
ROA spectra of viruses recorded in our laboratory have
displayed bands attributed to nucleic acid. However,
these are usually weak because of either the low nucleic
acid content of the virus or its flexible nature. Recently,
though, we have recorded the first nucleic acid ROA
spectra from an intact virus [67].
Cowpea mosaic virus (CPMV), the type member of
the comovirus genus, has a bipartite genome, meaning
that two RNA molecules (called RNA-1 and RNA-2)
are separately encapsidated. CPMV particles can be
separated into four distinct isolates by ultracentrifuga-
tion with a CsCl gradient [68]. The top component (T-
CPMV) consists of empty protein capsids, the middle
component (M-CPMV) contains capsids with a mole-
cule of RNA-2 bound, the bottom-upper component
(BU -CPMV) contains capsids with RNA-1 bound, and
the bottom-lower component (BL -CPMV) contains
capsids with RNA-2 bound plus Csþ ions permeating
the interior of the particles. The Raman and ROA
spectra of T-CPMV and M-CPMV are shown in Figs.
8a and b, respectively. CPMV has an icosahedral capsid
of T ¼ 3 symmetry, with two protein chains forming
alarge (L) coat protein and a third protein chain
forming a small (S) coat protein. Despite the low se-
quence homology between these three protein chains,
X-ray crystallography shows that they exhibit very
similar eight-stranded jelly roll b-barrel folds [68,69].
The details of the T-CPMV ROA spectrum are con-
sistent with those of a jelly roll b-barrel fold (see the
ROA spectrum of concanavalin A in Fig. 2b). There are Fig. 8. Backscattered Raman and ROA spectra of the icosahedral vi-
several large changes in the ROA spectrum of M- ruses (a) T-CPMV and (b) M-CPMV at pH 7.0 and 20 °C, and (c) the
CPMV due to new bands originating from the RNA-2 difference (b–a) (M-CPMV)-(T-CPMV).
molecule, which constitutes 24% of the particle by
mass. This is clearer in Fig. 8c which shows the Raman
and ROA spectra of the bound RNA-2 molecule ob- 5. Concluding remarks
tained by the subtraction (M-CPMV)-(T-CPMV).
Bands corresponding to conformational changes in- The sensitivity of ROA to molecular chirality makes
duced in the protein structure by nucleic acid binding it a valuable new tool for investigating biomolecular
may also be present. Comparison of this ROA spec- structure and behavior in solution which may now be
trum with those of the nucleic acids shown in Fig. 5 applied routinely to a wide range of proteins, nucleic
indicates that RNA-2 has an A-type helical conforma- acids, and viruses. It is already providing information
tion. A similar analysis has also shown that the RNA-1 complementary to that obtained from high-resolution
molecule in the BU -CPMV particle has the same con- techniques such as X-ray crystallography and NMR
formation. The RNA is not observed in the X-ray spectroscopy. Particularly promising applications for
crystal structure of CPMV [69]. ROA spectroscopy include high-throughput protein fold
208 E.W. Blanch et al. / Methods 29 (2003) 196–209

recognition in structural proteomics, the protein mis- [24] E.W. Blanch, A.F. Bell, L. Hecht, L.A. Day, L.D. Barron, J. Mol.
folding diseases, and structural virology. With the ex- Biol. 290 (1999) 1–7, doi:10.1006/jmbi.1999.2871.
[25] S.-H. Lee, S. Krimm, J. Raman Spectrosc. 29 (1998) 73–80.
pected availability of a commercial instrument in the [26] M. Tsuboi, M. Suzuki, S.A. Overman, G.J. Thomas Jr.,
near future, we hope that this brief review will encour- Biochemistry 39 (2000) 2677–2684.
age wide use of ROA in biomolecular science. [27] P. Hanson, D.J. Anderson, G. Martinez, G. Millhauser, F.
Formaggio, M. Crisma, C. Toniolo, C. Vita, Mol. Phys. 95 (1998)
957–966.
[28] K.A. Bolin, G.L. Millhauser, Accs. Chem. Res. 32 (1999) 1027–
Acknowledgment 1033.
[29] G. Wilson, L. Hecht, L.D. Barron, J. Chem. Soc. Faraday Trans.
L.D.B. and L.H. thank the BBSRC for a research 92 (1996) 1503–1510.
grant. [30] A.A. Adzhubei, M.J.E. Sternberg, J. Mol. Biol. 229 (1993) 472–
493, doi:10.1006/jmbi.1993.1047.
[31] B.T. Stapley, T.P. Creamer, Prot. Sci. 8 (1999) 587–595.
[32] T. Miura, H. Takeuchi, I. Harada, J. Raman Spectrosc. 20 (1989)
References 667–671.
[33] E.W. Blanch, L. Hecht, L.A. Day, D.M. Pederson, L.D. Barron,
[1] P.R. Carey, Biochemical Applications of Raman and Resonance J. Am. Chem. Soc. 123 (2001) 4863–4864.
Raman Spectroscopies, Academic Press, New York, 1982. [34] E.W. Blanch, D.J. Robinson, L. Hecht, C.D. Syme, K. Nielsen,
[2] M. Diem, Modern Vibrational Spectroscopy, Wiley, New York, L.D. Barron, J. Gen. Virol. 83 (2002) 241–246.
1993. [35] L.D. Barron, E.W. Blanch, L. Hecht, Adv. Prot. Chem. 62 (2002)
[3] T. Miura, G.J. Thomas Jr., in: B.B. Biswas, S. Roy (Eds.), 51–90.
Subcellular Biochemistry Proteins: Structure Function and Engi- [36] O.B. Ptitsyn, Adv. Prot. Chem. 47 (1995) 83–229.
neering, 24, Plenum Press, New York, 1995, pp. 55–99. [37] M. Arai, K. Kuwajima, Adv. Prot. Chem. 53 (2000) 209–282.
[4] T.A. Keiderling, in: G.D. Fasman (Ed.), Circular Dichroism and [38] P. Haezebrouck, M. Joniau, H. vanDael, S.D. Hooke, N.D.
the Conformational Analysis of Biomolecules, Plenum Press, New Woodruff, C.M. Dobson, J. Mol. Biol. 246 (1995) 382–387,
York, 1996, pp. 555–598. doi:10.1006/jmbi.1994.0093.
[5] P.L. Polavarapu, Vibrational Spectra: Principles and Applications [39] L.A. Morozova-Roche, J. Zurdo, A. Spencer, W. Noppe, V.
with Emphasis on Optical Activity, Elsevier, Amsterdam, 1998. Receveur, D.B. Archer, M. Joniau, C.M. Dobson, J. Struct. Biol.
[6] G.J. Thomas Jr., Annu. Rev. Biophys. Biomol. Struct. 28 (1999) 130 (2000) 339–351, doi:10.1006/jsbi.2000.4264.
1–27. [40] E.W. Blanch, L.A. Morozova-Roche, D.A.E. Cochran, A.J. Doig,
[7] T.A. Keiderling, in: N. Berova, K. Nakanishi, R.W. Woody L. Hecht, L.D. Barron, J. Mol. Biol. 301 (2000) 553–563,
(Eds.), Circular Dichroism: Principles and Applications, second doi:10.1006/jmbi.2000.3981.
ed., Wiley, New York, 2000, pp. 621–666. [41] M.R.H. Krebs, D.K. Wilkins, E.W. Chung, M.C. Pitkeathly,
[8] B.R. Singh (Ed.), Infrared Analysis of Peptides and Proteins: A.K. Chamberlain, J. Zurdo, C.V. Robinson, C.M. Dobson, J.
Principles and Applications, in: ACS Symposium Series, vol. 750, Mol. Biol. 300 (2000) 541–549, doi:10.1006/jmbi.2000.3862.
American Chemical Society, Washington DC, 2000. [42] C.D. Syme, E.W. Blanch, C. Holt, R. Jakes, M. Goedert, L.
[9] L.A. Nafie, T.B. Freedman, in: N. Berova, K. Nakanishi, R.W. Hecht, L.D. Barron, Eur. J. Biochem. 269 (2002) 148–156.
Woody (Eds.), Circular Dichroism Principles and Applications, [43] C. Holt, L. Sawyer, J. Chem. Soc. Faraday Trans. 89 (1993) 2683–
second ed., Wiley, New York, 2000, pp. 97–131. 2692.
[10] L.D. Barron, L. Hecht, in: N. Berova, K. Nakanishi, R.W. [44] R.W. Carrell, D.A. Lomas, Lancet 350 (1997) 134–138.
Woody (Eds.), Circular Dichroism. Principles and Applications, [45] C.M. Dobson, R.J. Ellis, A.R. Fersht (Eds.), Philos. Trans. R.
second ed., Wiley, New York, 2000, pp. 667–701. Soc. Lond. B, 356, 2001, pp. 127–227.
[11] L.D. Barron, M.P. Bogaard, A.D. Buckingham, J. Am. Chem. [46] Y.C. Sekharudu, M. Sundaralingham, in: E. Westhof (Ed.), Water
Soc. 95 (1973) 603–605. and Biological Macromolecules, CRC Press, Boca Raton, FL,
[12] L.D. Barron, L. Hecht, E.W. Blanch, A.F. Bell, Prog. Biophys. 1993, pp. 148–162.
Mol. Biol. 73 (2000) 1–49. [47] G. Siligardi, A.F. Drake, Biopolymers 37 (1995) 281–292.
[13] P.W. Atkins, L.D. Barron, Mol. Phys. 16 (1969) 453–466. [48] M.L. Tiffany, S. Krimm, Biopolymers 6 (1968) 1379–1382.
[14] L.D. Barron, A.D. Buckingham, Mol. Phys. 20 (1971) 1111–1119. [49] R.W. Woody, Adv. Biophys. Chem. 2 (1992) 37–79.
[15] A.D. Buckingham, Adv. Chem. Phys. 12 (1967) 107–142. [50] R.K. Dukor, T.A. Keiderling, Biopolymers 31 (1991) 1747–
[16] L.D. Barron, Molecular Light Scattering and Optical Activity, 1761.
Cambridge University Press, Cambridge, 1982. [51] T.A. Keiderling, R.A.G.D. Silva, G. Yoder, R.K. Dukor, Bioorg.
[17] L.D. Barron, A.D. Buckingham, J. Am. Chem. Soc. 96 (1974) Med. Chem. 7 (1999) 133–141.
4769–4773. [52] N. Sreerama, R.W. Woody, Biochemistry 33 (1994) 10022–10025.
[18] L. Hecht, L.D. Barron, E.W. Blanch, A.F. Bell, L.A. Day, J. [53] A.F. Bell, L. Hecht, L.D. Barron, J. Chem. Soc. Faraday Trans.
Raman Spectrosc. 30 (1999) 815–825. 93 (1997) 553–562.
[19] W. Hug, G. Hangartner, J. Raman Spectrosc. 30 (1999) 841–852. [54] A.F. Bell, L. Hecht, L.D. Barron, J. Am. Chem. Soc. 119 (1997)
[20] A.T. Tu, Adv. Spectrosc. 13 (1986) 47–112. 6006–6013.
[21] S.J. Ford, Z.Q. Wen, L. Hecht, L.D. Barron, Biopolymers 34 [55] A.F. Bell, L. Hecht, L.D. Barron, Biospectroscopy 4 (1998) 107–
(1994) 303–313. 111.
[22] E.W. Blanch, L.A. Morozova-Roche, L. Hecht, W. Noppe, L.D. [56] A.F. Bell, L. Hecht, L.D. Barron, J. Raman Spectrosc. 30 (1999)
Barron, Biopolymers 57 (2000) 235–248. 651–656.
[23] L.D. Barron, E.W. Blanch, A.F. Bell, C.D. Syme, L. Hecht, L.A. [57] A.F. Bell, L. Hecht, L.D. Barron, J. Am. Chem. Soc. 120 (1998)
Day, It is my book, in: Hicks, J.M. (Ed.), Chirality: Physical 5820–5821.
Chemistry. ACS Symposium Series, vol. 810, American Chemical [58] W. Saenger, Principles of Nucleic Acid Structure, Springer, New
Society, Washington DC, 2002, pp. 34–49. York, 1984.
E.W. Blanch et al. / Methods 29 (2003) 196–209 209

[59] W. Chiu, R.M. Burnett, R.L. Garcia, Eds, Structural Virology of [65] S.B. Larson, J. Day, A. Greenwood, A. McPherson, J. Mol. Biol.
Viruses, Oxford University Press, New York, 1997. 277 (1998) 37–59, doi:10.1006/jmbi.1997.1570.
[60] D.A. Marvin, Curr. Opin. Struct. Biol. 8 (1998) 150–158. [66] K. Valegard, L. Liljas, K. Fridborg, T. Unge, Nature 345 (1990)
[61] D.M. Pederson, L.C. Welsh, D.A. Marvin, M. Sampson, R.N. 36–41.
Perham, M. Yu, M.R. Slater, J. Mol. Biol. 309 (2001) 401–421, [67] E.W. Blanch, L. Hecht, C.D. Syme, V. Volpetti, G.P. Lomonoss-
doi:10.1006/jmbi.2001.4685. off, L.D. Barron, J. Gen. Virol. 83 (2002) 2593–2600.
[62] E.W. Blanch, D.J. Robinson, L. Hecht, L.D. Barron, J. Gen. [68] G.P. Lomonossoff, J.E. Johnson, Prog. Biophys. Mol. Biol. 55
Virol. 82 (2001) 1499–1502. (1991) 107–137.
[63] A. Klug, Philos. Trans. R. Soc. Lond. B 354 (1999) 531–535. [69] T. Lin, Z. Chen, R. Usha, C.V. Stauffacher, J.-B. Dai, T. Schmidt,
[64] G. Stubbs, Philos. Trans. R. Soc. Lond. B 354 (1999) 551– J.E. Johnson, Virology 265 (1999) 20–34, doi:10.1006/
557. viro.1999.0038.

Вам также может понравиться