Summary of VCD concepts and techniques

Bulk of text is taken from a review written by Tim Keiderling for

Infrared and Raman Spectroscopy of Biological Materials

ed. Bing Yan and H.-U. Gremlich, Marcel Dekker, New York (2001) p.55-100

VIBRATIONAL CIRCULAR DICHROISM OF PEPTIDES AND PROTEINS

Survey of Techniques, Qualitative and Quantitative Analyses, Applications

Vibrational Circular Dichroism (VCD, the differential absorption of left and right circularly polarized light or DA=AL -AR, by fundamental vibrational transitions of a chiral molecule) can be seen as a natural outgrowth of Electronic CD (ECD) technology, commonly measured for chiral molecules in the uv region, lending stereochemical sensitivity to broad featureless electronic absorption transitions. ECD originally developed for study of the absolute configuration, and conformation (if the configuration is established independently) of organic molecules. However the molecules needed accessible transitions, which limited the techniques to chromophoric molecules, with a big emphasis on C=O containing functionalities and the coupling of dipole transitions in polymeric systems. The latter fit the needs of biochemistry well, the amide linkage became a prime target of study in peptides and proteins and biochemical secondary structure analyses and monitoring (for example in protein folding) became the dominant application of ECD. By comparison, infrared (IR) and Raman spectra have a natural resolution advantage as compared to the electronic region of the spectrum, and the chromophores are the bonds of the molecules itself, offering ideal probes of stereochemistry. However, IR (and Raman) depend on the square of the dipole (or polarizability) change in the vibration, naturally leading to all transitions having a single sign (positive). VCD and the conceptually related Raman optical activity (ROA) techniques developed to give the stereochemical sensitiveity of chiroptical spectra this vibrational resolution advantage. In other words, VCD is to IR what ECD is to uv absorption spectra, nothing is lost but much is gained.

In the ensuing 25+years since their initial development, theory and applications for VCD and ROA have developed apace. This site contains information on our biopolymer applications of VCD as well for other techniques and problems. A brief discussion of ROA can be found by linking here and elsewhere in the extensive reviews of Laurence Barron.

Predictions of VCD based on simple classically based models were published before the first experiments appeared. The first experimental VCD was published by George Holzwarth (then at U. of Chicago) in conjunction with a theoretical interpretation by Al Moscowitz at Minnesota and their students. Following this Larry Nafie, Jack Cheng and Tim Keiderling worked in Philip Stephens lab at USC to develop an instrument that would measure improved VCD in the nearIR (e.g. C-H stretch) and demonstrate the applicability (range of molecules accessible) with the VCD technique, which are described in instrument and application papers from USC by those authors in 1975-77. Since then several groups have developed instrumentation, theory and applications for VCD, with the Stephens (USC), Nafie (Syracuse, and Keiderling (UIC) labs being active over the entire period and those of Polavarapu (Vanderbilt), Diem (Hunter, CUNY), Weiser (Calgary), Urbanova (Tech Univ, Prague), Malon and Bour (Acad. Sci., Prague), Hirota (Osaka), Asher (Pittsburgh), Drake (Kings Coll. London) contributing at various times. Several have worked in biomolecular applications, which form the remainder of this discussion. Now for the review:

1. Introduction

Determination of protein secondary structure has long been a major application of optical spectroscopic studies of biopolymers . In most cases these efforts have been aimed at evaluating the average fractional amount of helix and sheet contributions to the overall secondary structure in a protein or peptide. In some cases further interpretations in terms of turns and specific helix and sheet segment types have developed. This focus on average secondary structure is a consequence of the interactions of primary importance to these techniques and their relatively low resolution, which ordinarily does not provide site-specific information without technologically challenging selective isotopic substitution. Such a limit is in contrast to x-ray crystallography or nmr spectroscopy which naturally yield site-specific structural information, due to their very high resolution (at least for nmr), requiring at most uniform labeling. However, these invaluable structural biology techniques are very slow both in terms of data acquisition and in completion of the complex interpretive process. Furthermore, the intrinsic time scales of these measurements are slow, thereby not permitting reliable analysis of dynamic structures or of conformations undergoing fast changing events. Only more limited applications of optical spectra to determination of the tertiary structure or fold of the secondary structural elements have appeared, and these have typically used fluorescence or near-uv electronic circular dichroism (ECD) of aromatic residues to sense a change in the fold rather than to determine its nature .

For biomolecular structural studies, ECD of transitions in the ultraviolet has been one of the dominant applications of that technique due to its sensitivity to molecular conformation, which is often manifested in complete sign reversals for selected structural changes. Most spectroscopic secondary structure studies of proteins and peptides have used far-uv ECD for the n-p * and p -p * transitions of the amide linkage . Coupling of the involved electronic transition dipoles leads to the extended chirality which is characteristic of the chain conformation and results in the observed CD spectrum which has a sign/frequency profile that varies for different secondary structures. Since these transitions are broad and fully overlapped, their individual frequencies do not have much impact on the correlation of CD spectral properties with structural features. On the contrary, the usual interpretations employ bandshape-based schemes that are either qualitative in nature or dependent on a statistical fit to a set of (typically protein) spectra which provide the structural reference set . While the earliest methods utilized data from polypeptides for qualitative and quantitative analyses, more reliable results were obtained using a basis set of proteins whose crystal structures were known.

IR and Raman analyses of secondary structure historically took a different approach , due to the natural resolution of the vibrational region of the spectrum into contributions from modes characteristic of different bond types in the molecule. Initial focus was on assigning component frequencies to various secondary structural component types as has been discussed in several reviews . Most effort focussed on the mid IR amide I (C=O stretch) band with additional use of the amide II (N-H deformation plus C-N stretch) in IR spectra and amide III (oppositely phased N-H deformation plus C-N stretch) with Raman spectral methods. In non-aqueous media, the near IR amide A (N-H stretch) can also be useful. Since both IR and Raman techniques give rise to single-signed spectral bandshapes that are effectively just the dispersed sum of the contributions from all the component transitions, and those components differ by only relatively small amount in frequency, the bandshapes for different proteins are very similar. However, due to the high signal to noise ratio (S/N) of Fourier transform IR (FTIR), this approach can be pushed further with resolution enhancement using second derivative or Fourier self deconvolution (FSD) techniques . These latter methods partially compensate for the relatively low level of bandshape variation available in the IR and Raman spectra, but are subject to abuse by the unwary user and submit the analysis to several assumptions that may well be unwarranted . Most notably, the frequencies assigned to specific secondary structural types are assumed to be unique, whereas any given type will shift significantly under influence of different solvents and residues . Additionally non-uniformity and end effects for these segments of a given structural type will have significant impact on the frequencies resulting in some dispersion of the contribution of a given secondary structure segment over the spectrum . Finally most methods assume that the dipole strengths (extinction coefficients) of all the residues, regardless of their conformation, is the same, while they in fact vary . None-the-less such FTIR methods have proven to be very useful for sorting out differences between proteins with similar IR spectra. Bandshape-based analyses similar to those used for ECD have also been applied to FTIR and Raman spectra with reasonable success as seen with ECD .

This contrast of FTIR and ECD sensitivities has led to the development of vibrational (or infrared) CD (VCD, or DA=AL-AR, the differential absorbance for left and right circularly polarized light by vibrational transitions in IR) and its counterpart, Raman optical activity (ROA) in several laboratories. Only the former measurement will be addressed here; but reviews of ROA abound . The key impetus for moving to the vibrational region of the spectrum is that it is rich with resolved transitions, which are characteristic of localized parts of the molecule, and that by making a VCD measurement one can impart distinct stereochemical sensitivity to each of them . The chromophores needed in the molecule for VCD measurement are simply the bonds themselves as sampled by their stretching and bond deformation excitations. Chiral interaction of these bonds (necessary to establish optical activity) will then be manifest in the spectral bandshape that results from a VCD measurement. Furthermore these vibrational excitations are part of the ground state of the molecule (where all thermal processes occur, such as conformational variations) for which one normally wishes to gain structural insight. VCD has the three dimensional structural sensitivity natural to a chiroptical technique, but this is manifested in a response distributed over a large number of localized probes of the structure .

Of course this benefit comes at a cost, which arises from significantly reduced S/N and some theoretical interpretive difficulty as compared to IR. Developments on the latter front are fast bringing the theoretical capability for prediction of VCD spectra for small molecules to a level that is demonstrably superior to that for ECD spectra . However, these ab initio quantum mechanical methods are severely limited if one were to apply them to large molecules, such as proteins. Thus most biomolecular applications of VCD use empirically based analyses . Experimentally, instrumentation has reached a stage where VCD spectra for most molecular systems of interest can be measured under at least some sampling conditions . It is true that most VCD studies of biomolecules in aqueous solutions, naturally the conditions of prime interest, are restricted to relatively high concentration samples, much as is characteristic of nmr. (By comparison, ROA measurements demand even higher concentrations .)

VCD has the bandshape variability of CD coupled with the frequency resolution of IR which leads to a demonstrably enhanced sensitivity to secondary structure in proteins . Sampling conditions for obtaining experimental VCD spectra of protein and peptide samples are similar to those used in FTIR studies with the important exception that the data are differential spectra of much smaller amplitude (DA is the order of 10-4 - 10-5 of the sample absorbance, A) and thus lower S/N. Consequently, to obtain quality VCD spectra, much longer data collection times are required than for FTIR or ECD and specially designed instruments are used. Experimental methods of biomolecular VCD are summarized in the next section and are fully discussed in separate reviews . Theoretical techniques for simulation of small molecule VCD are also the focus of several previous reviews and will be only briefly surveyed in a section following that. To date, theory has played a minor role in the interpretation of VCD spectra of most biomolecules due to their large size. However, recent advances in computational techniques and hardware have made larger calculations on sizeable, at least realistic peptides, ever more feasible.

On the other hand, qualitative analyses can be done with facility utilizing the VCD bandshape and its frequency position to predict the dominant secondary structural type in a peptide or protein. This has become the standard approach for most peptide studies . In such smaller fully solvated, biopolymers, the frequency shifts due to solvent effects and the inhomogeneity of the peptide secondary structure (as well as end fraying) are severe problems for frequency based analyses. In globular proteins, qualitative estimations of structure remain of interest for determining the dominant fold type, eg. highly helical, highly sheet or mixed helix and sheet type, but quantitative estimations of secondary structure content based on empirical spectral analyses are usually of more interest. Quantitative methods for analysis of protein VCD spectra in terms of structure follow the lead of most ECD analyses and employ bandshape techniques referenced to a training set of protein spectra . In this respect, globular proteins in aqueous solution are assumed, on average, to have their peptide segments in similar environments and of similar lengths, so that the solvent and length effects on the peptide modes will be relatively consistent for the training set and any unknowns studied.

In summary, unlike ECD, VCD can be used to correlate data for several different spectrally resolved features; and, unlike IR and Raman spectroscopies, each of these features will have a physical dependence on stereochemistry. But from another point of view, the combination of these techniques can compensate for each other, providing a balance between accuracy and reliability. The prime questions remaining in the VCD field now relate to application and interpretation of the method. It is clear that, despite claims of fundamental advantages of any one technique, progress in understanding of biomolecular structures will come from synthesizing all the data gathered from various techniques. In our biomolecular work, different types of spectral data are used to place bounds on the reliability of structural inferences that might be drawn from any one technique. Furthermore, such insight into average secondary structure may provide constraints on site-specific structure analyses using nmr and computational methods.

2. Experimental Techniques

This new dimension in optical activity comes at some cost in that the rotational strengths of vibrational transitions as detected in VCD are much weaker than are those of electronic transitions detected in ECD. Similarly since VCD is a differential IR technique, its S/N can never approach that of FTIR which represents a summed response. Several research groups have developed instrumentation that makes the measurement of VCD reasonably routine over much of the IR region commercial FTIR vendors are now providing VCD accessories or, in one instance (Bomem-Biotools), a stand-alone VCD instrument that has now been shown by its users to have an exceptional S/N and baseline characteristics . In this section, instrument designs are briefly summarized and compared.

2.1 FT-VCD vs. DISPERSIVE VCD.

Available instrumentation makes routine measurement of VCD possible over much of the IR region down to ~700 cm-1 on many samples. Development of a VCD instrument is normally accomplished by extending a dispersive IR or an FTIR spectrometer to accommodate, in terms of optics, time-varying modulation of the polarization state of the light and, in terms of electronics, detection of the modulated intensity that results from a sample with non-zero VCD. Our instruments and those of others are described in the literature in detail as referenced in recent reviews . A detailed review contrasting these designs and detailing components needed to construct either type of instrument has been published by this author . Here only a brief survey of the important components is given.

VCD instruments share several generic elements with "normal" CD instruments. All current instruments use a broad band source of light, typically utilizing black-body radiation from something like a ceramic or graphite-based glower (or tungsten in the near IR), to allow sampling of a spectrum over the IR region. The method chosen for encoding the optical frequencies divides VCD instruments into two styles. Dispersive VCD instruments use a monochromator, based on grating technology, which scans only through the wavelength spectrum of interest, recording the response sequentially. Such an instrument must be optimized for efficient light collection. On the other hand, Fourier transform (FT) VCD instruments use a Michelson interferometer that encodes the optical frequencies as an interferogram and gains efficiency through the multiplex advantage. With FT-VCD, spectral responses at all wavenumbers of interest are obtained simultaneously, with longer scans serving to improve the resolution of spectral features. Both styles of instrument then are modified to provide for linear polarization of the light beam, normally with a wire grid polarizer, and modulation of it, with a photo-elastic modulator (PEM), between (elliptically) right- and left-hand polarization states. The beam then passes through the sample and onto the detector, typically a liquid N2 cooled Hg1-x(Cd)xTe (MCT) photoconducting diode. After preamplification, the electrical signal developed in the MCT is separated to measure the overall transmission spectrum (Itrans) of the instrument and sample via one channel and the polarization modulation intensity (Imod), which is related to the VCD intensity, via the other.

These signals are ratioed to yield the raw VCD signal either before or after A-to-D conversion, depending on the instrumental design. Since VCD is a differential absorbance measurement, D A=AL-AR for left/right (L/R) circularly polarized light, it is necessary to ratio these two intensities to normalize out any dependence on the source intensity and instrument transmission characteristics. In the limit of small D A values,

Imod/Itrans = (1.15 DA)J1(a0)gI (1)

where J1 (a0) is the first order Bessel function at the maximum retardation of the modulator, a0, and gI is an instrument gain factor. Evaluation of this term and elimination of the gain factor are obtained by calibration of the VCD using a pseudo sample composed of a birefringent plate and a polarizer pair or by measuring the VCD of a known sample . Further processing of the computer stored VCD spectrum involving calibration, baseline correction and spectral averaging or smoothing, as desired, and conversion to molar quantities, e.g. D e = D A/bc where b is the path length in cm and c is the concentration in moles/L, completes the process. To give these concepts more substance, some details of the UIC dispersive and FTIR based instruments are given below.

Insert dispersive Instrument slide

Our original dispersive instrument is configured around a 1.0 m focal length, ~f/7 monochromator (Jobin-Yvon, ISA) that is illuminated with a home built carbon rod source . A mechanical chopper provides the modulation necessary for detecting the instrument transmission with an MCT detector. The monochromator output is filtered with a long-wave pass interference filter (OCLI) to eliminate light due to higher order diffraction from the grating and uses mirrors (achromatic) to focus the beam on the sample. A more recent, compact design has been shown to have advantages in terms of S/N and baseline stability .

The light is linearly polarized by means of a wire grid polarizer on a BaF2 substrate (Cambridge Physical Sciences, Molectron Detectors in the USA) and modulated between left and right circularly polarized states with an AR coated ZnSe PEM (Hinds Instruments). Following the sample a ZnSe lens focuses the light onto an MCT detector chosen in terms of size and shape to match the slit image (we have an array of 3 2mm x 2mm elements) for optimal operation down to ~800 cm-1. Alternatively, very high sensitivity in the near-IR (~5000-1900 cm-1) is possible with an InSb photovoltaic detector, CaF2 modulator and lens, and a grating optimized for that region. Lower frequencies can be accessed with different detectors but with loss of S/N .

To process the signal, a lock-in amplifier is used to detect the transmission intensity of the instrument and sample as evidenced by the signal developed in phase with the chopping frequency. The polarization modulation intensity is measured with a separate lock-in amplifier as that component of the detector signal in-phase with the PEM frequency. Since the VCD is also modulated by the chopper, the signal can be demodulated again by using another lock-in referenced to the chopper, but now measuring the output of the one referenced to the PEM. Dynamic normalization varies the amplification gain such that the transmission signal is constant. Applying the same gain to the polarization modulated signal assures a normalization much as is accomplished in "normal" CD instruments (which vary the high voltage applied to the photomultiplier to vary its gain and yield a constant average, or dc, current). In an alternate design, both signals can be A-to-D converted and normalization effected by digital division in the data computer .

Insert slide of FT-VCD

Our FTIR-VCD spectrometer uses a Digilab (BIORAD) FTS-60A FTIR as its core but the choice of FTIR is fully open to the user, since these optics seem to impart little limitation for practical VCD operation and very successful instruments in other laboratories have been configured around a number of FTIRs . In a separate compartment, an external beam goes through the same type of linear polarizer, stress optic modulator, lens and relatively large area MCT detector as described above, with only weak focussing at the sample which leads to flatter baselines. MCT detectors can easily saturate to a non-linear response due to the high light levels in an FTIR. This can be controlled by using optical filters (e.g. 1900 cm-1 cut-off low-pass) to isolate the spectral region of interest and by controlling the preamp gain. For aqueous, biological samples, the spectral band pass is limited by the solvent, so high light level is not a major problem.

The raw VCD is obtained by ratioing the spectrum of the polarization modulated signal with the normally developed transmission single beam spectrum using the FTIR computer processing software. In a rapid scan instrument, the detector signal is processed by a lock-in amplifier referenced to the modulator to form an ordinary interferogram of the modulated signal which the FTIR electronics can process. Most rapid scan FTIR VCD spectra have concentrated on mid-IR bands since the near-IR corresponds to higher frequency sidebands, which the lock-in attenuates. Slow- or step-scan operation yields better response for higher frequency, near IR, components of the spectrum . The optical frequencies are also encoded through correlation to the mirror position, as measured by use of laser fringe counting, but the time element is removed.

While the instrument throughput (equivalent to Itrans) is an ordinary IR intensity measurement for which all instruments are adequately programmed to process, the polarization modulation signal (Imod) is not so simply processed. Often the integral of the modulated spectrum is very small having a rough balance between positive and negative VCD bands. This results in their being only a very weak center burst in the interferogram. For purposes of interferometer alignment and phase correction, this can pose difficulties . Normally, specialized software is required to overcome these limits, provide for simultaneous or sequential measurement from two independent detector (Itrans and Imod) inputs, and permit a variety of arithmetic manipulations of the Itrans and Imod spectra. Unlike the case for FTIR hardware, software is a central consideration in choosing an instrument for VCD use.

While FT-VCD has many advantages, the restriction to measurement only in the spectral windows of water and the relatively broad bands seen in biopolymer IR spectra can nullify the multiplex and throughput advantages of FTIR and, all other things being equal, favor use of dispersive VCD. Until now, the expected FTIR advantages have not been experimentally realized in terms of the S/N for low resolution biomolecular (aqueous) FTIR-VCD spectra as compared to what can be measured with the dispersive instrument over a similar time span FT-VCD measurement, which always encompasses the full spectrum, can be quite inefficient as compared to concentrating oneís effort and maximizing S/N in just one or two spectral bands, such as are accessible for a protein or peptide in D2O.

Insert slides comparing dispersive and ChiralIR or Dispersive and our FT-VCD

To get adequate signal to noise ratio (S/N) and determine scan-to-scan reproducibility, the dispersive spectra are averaged for several scans, often using time constants of the order of 10 sec and resolutions of ~10 cm-1. This means a typical IR band can take about 1/2 hour for a single scan. FTIR measured VCD spectra can sample a much wider spectral region and take ~1/2 hour (on our instruments) to collect an adequate number of scans for detecting the features of interest at higher resolution for a rigid, chiral organic molecule, but would require extensive averaging over much longer times to match the S/N available using the dispersive instrument for single bands in aqueous phase biopolymers. If, in the end, only one or two adjacent bands are needed for the analysis, much time can be lost with the FTIR-based technique; but if multiple bands are to be studied, FTIR-VCD retains its advantage, even for biological samples . In both cases, these inherently single-beam, though corrected for light-beam intensity by the Imod/Itrans ratioing step, VCD scans must be coupled with equally long collections of baseline spectra to correct for instrument and sample induced spectral response. Finally in this comparison of FTIR and dispersive based VCD instruments, it might be noted that the more direct dispersive measurement is intuitively easier to interpret in case something goes wrong, such as noise or baseline artifact.

Due to its weak signal size, VCD is subject to artifacts which must be corrected by careful baseline subtraction. The best baseline is determined using racemic material, which is impractical for most biological materials. However, satisfactory baselines for spectral corrections can often be acquired with carefully aligned instruments by measuring VCD spectra of the same sample cell filled with just solvent. There exists no satisfactory theory of these artifacts that can be used to control baselines. Rather there is an empirical body of evidence that parallel or slowly converging beams, few reflections and uniform detector surfaces give the best results. (More recently, Nafie, in Applied Spectr. 2000, has shown that use of a double modulator approach can largely correct baseline offset and eliminate much of the worst artifacts.) Finally, most artifacts can also be minimized by careful optical adjustments which, at least in our instruments, are very stable, not requiring corrections for months. In this case, baseline subtraction also subtracts the bulk of the artifacts.

2.2 SAMPLING TECHNIQUES.

Most biomolecular systems are best studied in an aqueous environment. This poses difficulties for IR techniques due to water being a very strong absorber whose fundamental transitions strongly overlap regions of interest in biomolecules such as the N-H and C=O stretches. Consequently, the peptide amide I' (primes for N-deuterated amides) band at ~1650 cm-1, which is dominated by the C=O stretch, has normally been measured in D2O based solution. On the other hand, the amide II at ~1550 cm-1 and amide III at ~1300 cm-1 are best studied in H2O.

Protein samples in D2O can be prepared at concentrations in the range of 20-50 mg/ml for VCD. An aliquot of the solution (typically 20-30 ml) is placed in a standard demountable cell consisting of two BaF2 windows separated by a 25-50 µm Teflon spacer. For studies in H2O, concentrations of >100 mg/ml (but <20 ml in volume) and pathlengths of 6µm are most useful to allow detection of amide I, where the water alone gives absorbances of ~0.9 at 1650 cm-1. This interference causes loss in S/N, but no major artifacts are found in our instruments when properly aligned. For these experiments we find it most useful to use refillable cells (Specac) and to run the H2O baseline first, then replace the solvent with sample without demounting .

Insert Slide of demountable cell

Final VCD curves are obtained by subtraction of a baseline VCD scan from the sample spectrum and by calibration as noted above. Typically, after obtaining baseline and sample VCD scans, single beam IR transmission spectra of the sample and of the solvent are recorded in the same cell to obtain an absorbance spectrum, ideally this should be done on the same instrument and under identical conditions as were the VCD spectra. It is often useful to obtain FTIR spectra at higher resolution and optimal S/N on the same samples for purposes of comparison and for resolution enhancement of the absorption spectrum using Fourier self-deconvolution . The FTIR spectra can also be used to frequency correct the dispersive VCD spectra. Ideally, VCD should be plotted in molar units such as e and D e as is done commonly with ECD measurements. However, since concentration and path lengths are rarely known to sufficient accuracy, VCD spectra of biomolecules are often normalized to the absorbance which in this case should be measured on the VCD instrument for consistency. Because the absorbance coefficients for different molecules studied will vary, this is only a first order correction for concentration.

In our laboratory, ECD spectra are additionally measured for the samples studied using a commercial instrument (now a JASCO J-810). These spectra are usually obtained under more dilute conditions using strain-free quartz cells (NSG Precision Cells or Helma Cells) obtained with various sample path lengths from 0.1 to 10 mm, the shorter path length cells being somewhat difficult to clean. Since relatively small amounts of biopolymer can give rise to significant ECD signals, it is very important to thoroughly clean sample cells between uses. Concentrations used in our laboratory for ECD are often of the order of magnitude of 0.1-1 mg/ml. For comparison of data obtained under comparable conditions, it is possible to measure ECD on the same samples used for VCD (> 10 mg/ml) by employing 6-15 µm path cells constructed with quartz windows and a teflon spacer .

References: