Fact-checked by Grok 2 weeks ago

Cotton effect

The Cotton effect is the characteristic anomalous change in (ORD) and/or (CD) observed in optically active substances near an electronic absorption band of a chiral , typically appearing as a maximum or minimum in the ORD curve accompanied by a sign inversion. This phenomenon, which reflects the differential interaction of circularly polarized light with enantiomeric forms, was first discovered and described in 1895 by French physicist Aimé Cotton during his studies on the rotation of light in absorbing media. The physical basis of the Cotton effect stems from the unequal velocities and absorption rates of left- and right-circularly polarized light components within a chiral , particularly when the incident approaches an such as an n→π* or π→π* in chromophores like carbonyl groups around 290 . In ORD spectra, a positive Cotton effect is defined by an initial rise to a positive extremum followed by a crossover to negative rotation, as seen in rigid trans-decalone derivatives, whereas a negative effect exhibits the inverse pattern, often in cis isomers. Similarly, CD measurements reveal corresponding positive or negative bands, quantifying the intensity of this differential absorption. Historically, early investigations in the focused on carbonyl-containing compounds like steroids and terpenoids to correlate Cotton effect signs with molecular configuration, building on Cotton's foundational work. Today, the Cotton effect serves as a critical tool in chiroptical for assigning absolute , analyzing enantiomeric purity, and studying conformational in biomolecules such as proteins and nucleic acids, where secondary structures produce distinct signatures.

Overview

Definition

The Cotton effect is a fundamental phenomenon in chiroptical , characterized by a distinctive alteration in the (ORD) and/or (CD) spectra of chiral substances near an absorption band. In ORD, it manifests as a rapid variation in the of plane-polarized light with wavelength, typically exhibiting an S-shaped curve with a peak and a trough flanking the absorption maximum, where the rotation often crosses zero. This anomaly arises because the absorption of circularly polarized light components influences the overall rotation, with the effect being either positive (rotation increases to a maximum before decreasing through zero) or negative (the opposite sequence). In terms of , the Cotton effect appears as an asymmetric absorption band, where one of circularly polarized light is absorbed more strongly than the other in the vicinity of the electronic transition, leading to a positive or negative signal that correlates directly with the ORD signature. This differential absorption is quantified by the difference in molar extinction coefficients for left- and right-circularly polarized light, underscoring the Cotton effect's role in revealing the chiral environment's impact on light-matter interactions. The phenomenon is observed exclusively in optically active compounds, such as those with atoms or helical structures, where the proximity to an absorption band amplifies the chiroptical response. Overall, the Cotton effect encapsulates the interplay between light and optical activity, providing a spectral fingerprint for that is most pronounced when the approaches the band's center, often resulting in a zero-crossing of at or near the . This defines its utility in distinguishing conformational and configurational aspects of chiral molecules through measurable deviations in rotatory power.

Importance

The Cotton effect serves as a key tool for studying molecular and conformational analysis, providing sensitive insights into the three-dimensional arrangement of chiral molecules through characteristic anomalies in (ORD) and (CD) spectra. This phenomenon allows researchers to probe the and dynamic conformational changes in , offering a direct correlation between spectral features and stereochemical properties without relying on invasive techniques. One significant advantage of the Cotton effect is its role in enabling non-destructive determination of stereochemical features, as it involves spectroscopic measurements on solutions that preserve sample integrity and do not require , unlike methods. This approach is particularly valuable for analyzing delicate or limited quantities of chiral compounds, facilitating rapid and reversible assessments of enantiomeric purity and structural dynamics. In biomolecular research, the Cotton effect is fundamental for understanding structures such as protein secondary elements, where β-sheets exhibit a characteristic negative Cotton effect around 217 nm due to the n–π* transition, aiding in the identification of folding patterns and interactions in native environments. By bridging and , it influences diverse fields, including for monitoring asymmetric transformations and pharmaceuticals for ensuring the stereochemical integrity of bioactive molecules.

History

Discovery

The Cotton effect was discovered in 1895 by French physicist Aimé Cotton (1869–1951) during his investigations into the of colored tartrates. While examining solutions of these chiral compounds, Cotton observed an anomalous dispersion in near absorption bands, marking the first identification of this phenomenon in absorbing media. Cotton first detailed these findings in his doctoral thesis, published in the Annales de Chimie et de Physique in 1896, where he reported measurements of anomalous rotation specifically in copper tartrate complexes. His experiments involved preparing optically active solutions of copper(II) tartrate, derived from , and recording the rotation of plane-polarized light as a function of close to the visible around 600 nm. These observations revealed a characteristic reversal in the sign of rotation at the absorption maximum, contrasting with the smooth seen in non-absorbing regions. The phenomenon was named the Cotton effect in recognition of his pioneering contributions to the study of optical activity in absorbing media, establishing a foundational concept in chiroptical . Cotton's initial work focused on quantitative measurements of both and differential absorption of circularly polarized light in these solutions, laying the groundwork for subsequent in the field.

Early Developments

Following Aimé Cotton's seminal 1895 publication on the anomalous rotatory dispersion near absorption bands, subsequent research in the late 1890s and early 1900s explored related magneto-optical phenomena. In 1907, Cotton collaborated with Henri Mouton to investigate magnetic in liquids, leading to the discovery of the Cotton-Mouton effect, which describes the induction of by a transverse and is distinct from the chiral Cotton effect observed in optically active substances. In the post-1895 period, chemists began applying the Cotton effect to inorganic coordination complexes to probe . , the founder of coordination chemistry, utilized (ORD) measurements in the early 1910s to support his octahedral configuration theory, notably resolving and analyzing the ORD of chiral cobalt(III) complexes like [Co(en)3]Cl3, where en is ; these studies demonstrated anomalous dispersion consistent with the Cotton effect near ligand absorption bands, providing key evidence for in metal complexes. The 1920s and marked significant advancements in ORD instrumentation and its extension to compounds, enabling more precise measurements of Cotton effects. Pioneers such as developed methods to study rotatory dispersion in absorbing media, publishing detailed analyses of Cotton effects in molecules like derivatives in the early ; similarly, Waldemar Kuhn and Samuel Mitchell contributed foundational work on ORD curves for chiral s, with improvements in visual and early photoelectric polarimeters allowing wavelength-dependent scans into the UV region for better resolution of anomalous dispersion extrema. In the , the development of improved photoelectric polarimeters revitalized ORD studies, particularly for . Chemist and collaborators extensively applied ORD to determine configurations in products such as steroids, correlating Cotton effect signs with molecular structure and establishing the method's value in .

Fundamentals

Optical Rotatory Dispersion

Optical rotatory dispersion (ORD) refers to the variation in the optical rotation of plane-polarized light as a function of wavelength when passing through a chiral medium. This phenomenon arises from the differential refractive indices for left- and right-circularly polarized light in optically active substances, leading to a rotation of the polarization plane that depends on the wavelength. In regions far from electronic absorption bands, ORD exhibits normal dispersion, characterized by a gradual, monotonic change in the rotation angle with , often following a smooth curve without abrupt variations. However, near an absorption band of the chiral molecule, the dispersion becomes anomalous, resulting in a sharp, non-monotonic variation known as the Cotton effect. This anomalous ORD manifests as an S-shaped curve in the plot of [\alpha] versus \lambda, where the rotation increases or decreases rapidly through the absorption region, crossing zero at a point often corresponding to the absorption maximum. ORD measurements are typically conducted using a equipped with a wavelength-tunable , where the [\alpha] is recorded as a function of \lambda across the spectral range of interest. The S-shaped dispersive curve in anomalous regions provides a signature of how electronic absorption perturbs the optical rotation in chiral media, serving as a foundational concept for interpreting related chiroptical phenomena.

Circular Dichroism

Circular dichroism (CD) is a spectroscopic technique that measures the differential of left-circularly polarized (LCP) and right-circularly polarized (RCP) by chiral molecules, quantified as Δε = ε_L - ε_R, where ε_L and ε_R are the molar absorptivities of LCP and RCP , respectively. This difference arises near bands of chromophores in asymmetric environments, where the chiral center induces unequal of the two polarizations. In CD spectra, peaks or troughs appear at wavelengths corresponding to the electronic transitions of absorbing groups, with the sign (positive or negative) and magnitude of these features revealing the absolute configuration and handedness of the chiral structure. For enantiomers, the CD signals are equal in magnitude but opposite in sign, providing a direct probe of molecular chirality. CD is frequently used alongside optical rotatory dispersion (ORD) to study chiroptical properties, where the Cotton effect manifests as characteristic extrema in both spectra near absorption bands. This complementary approach enhances the interpretation of chiral phenomena by combining absorption differences from CD with rotation changes from ORD. CD data are typically reported in units of molar ellipticity [θ] (in deg·cm²·dmol⁻¹) or Δε (in M⁻¹·cm⁻¹) and plotted as a function of , often in the ultraviolet-visible region to align with common absorptions.

Theoretical Basis

Physical Principles

Chiral molecules, lacking planes or centers of , exhibit a differential response to left-circularly polarized (LCP) and right-circularly polarized (RCP) light due to their intrinsic , which forms the basis of the Cotton effect. This asymmetry causes LCP and RCP light to experience distinct propagation speeds and absorption rates within the medium, leading to phenomena such as (CD) and (ORD). The effect is tied to the molecule's ability to distinguish between the helical wavefronts of the two polarizations, akin to a screw fitting a matching nut. Near absorption bands, the refractive indices for LCP and RCP light diverge markedly, producing anomalous dispersion where the optical rotation changes sign. This refractive index difference arises from the chiral medium's selective perturbation of the light's phase velocity, resulting in a characteristic S-shaped curve in ORD spectra centered around the absorption wavelength. The anomalous behavior is most evident close to electronic transitions, where the molecule's response amplifies the polarization-dependent scattering. At the quantum mechanical level, the Cotton effect originates from the coupling between electric dipole (μ) and (m) transition moments, which mix the with nearby excited states in chiral systems. This mixing, quantified by the rotational strength R = Im(μ · m), enables the otherwise forbidden magnetic dipole contributions to influence light-matter interactions, particularly when the electric and magnetic vectors are non-orthogonal due to molecular asymmetry. The effect is especially pronounced in the ultraviolet-visible (UV-Vis) spectral region for organic chromophores like carbonyl groups, where strong n→π* and π→π* electronic transitions occur around 200–300 nm, making these bands highly sensitive to chiral perturbations. In ketones, for instance, the weak inherent electric transition moment of the n→π* band at ~290 nm becomes CD-active through interactions with asymmetric neighboring groups.

Mathematical Formulation

The mathematical formulation of the Cotton effect centers on the rotational strength, which quantifies the magnitude and sign of the chiroptical response in (ORD) and (CD). The foundational expression for the rotational strength R_{0a} associated with a transition from the |0\rangle to an excited state |a\rangle is given by the Rosenfeld : R_{0a} = \Im \left[ \langle 0 | \boldsymbol{\mu} | a \rangle \cdot \langle a | \boldsymbol{m} | 0 \rangle \right], where \Im denotes the imaginary part, \boldsymbol{\mu} is the electric dipole transition moment operator, and \boldsymbol{m} is the magnetic dipole transition moment operator. This equation arises from quantum mechanical perturbation theory and links the Cotton effect directly to the scalar product of the transition moments, with the sign of R_{0a} determining whether the effect is positive or negative based on the relative orientation of these moments with respect to the molecular chiral center. The CD spectrum, which exhibits the Cotton effect as differential absorption, is often approximated by a Gaussian function to model the band shape near an absorption maximum. For a single band, the molar circular dichroism \Delta \epsilon(\sigma) as a function of wavenumber \sigma is expressed as: \Delta \epsilon(\sigma) = \Delta \epsilon_{\max} \exp\left[ -\frac{(\sigma - \sigma_0)^2}{2 \Delta \sigma^2} \right], where \Delta \epsilon_{\max} is the peak intensity, \sigma_0 is the central wavenumber, and \Delta \sigma characterizes the bandwidth. This Gaussian form provides a simple yet effective representation for simulating and analyzing the symmetric shape of isolated Cotton effects in CD spectra. ORD and CD are interconnected through dispersion relations, allowing one to be derived from the other via the Kramers-Kronig transform. The specific rotation [\alpha](\sigma) at wavenumber \sigma can be obtained from the spectrum as: [\alpha](\sigma) = \frac{1}{\pi} \mathcal{P} \int_0^\infty \frac{\Delta \epsilon(\sigma')}{\sigma'^2 - \sigma^2} \, d\sigma', where \mathcal{P} indicates the of the integral and \sigma is the in cm⁻¹. This transform captures how the anomalous in ORD, characteristic of the Cotton effect, emerges from the absorptive CD profile across all wavenumbers.

Characteristics

Positive and Negative Effects

The Cotton effect manifests in (ORD) spectra as anomalous dispersion near an band of an optically active , and it is classified as positive or negative based on the characteristic shape of the ORD curve. A positive Cotton effect is defined by a minimum in the ORD curve at the longer extremity of the band, followed by a maximum at the shorter wavelength side; this corresponds to an increase in the as the wavelength decreases through the absorption region. In contrast, a negative Cotton effect displays the opposite pattern, with a maximum at the longer and a minimum at the shorter , indicating a decrease in the as the decreases. This sign distinction arises from the interaction between the chiral environment and the electronic transitions of the , where the direction of the anomalous rotation reflects the of the molecular . The sign of the Cotton effect is intrinsically linked to the of the stereogenic center, providing a diagnostic tool for configurational ; for instance, trans-10-methyl-2-decalone exhibits a positive Cotton effect associated with the n→π* transition of the near 290 nm. In chiral derivatives, the octant rule predicts the sign based on the spatial arrangement of substituents relative to the carbonyl, with rear octants contributing oppositely to the rotational strength. The magnitude of the effect varies with the nature of the , being typically weaker for achiral chromophores (such as ) that are perturbed by adjacent chiral moieties, as the induced rotational strength depends on the degree of coupling between the electric and magnetic transition moments.

Spectral Curves

The (ORD) curve exhibiting the Cotton effect displays a characteristic S-shaped anomaly, where the changes sign with a zero-crossing point typically located near the of the corresponding (UV) absorption peak. This dispersive feature arises in the region of an electronic transition, contrasting with the relatively flat or gradually varying rotation outside absorption bands. In () spectra, the Cotton effect manifests as either a bisignate signal—with a positive and negative extremum—or a single extremum, directly corresponding to the differential of circularly polarized light during electronic transitions such as π→π* or n→π*. These features are absorptive in nature and align closely with the UV , often showing extrema at wavelengths where the ORD zero-crossing occurs. Such spectral anomalies commonly appear in the UV region for π→π* transitions around 200-250 and in the near-UV for n→π* transitions around 280-300 , particularly for carbonyl-containing chromophores. For example, (+)-3-methylcyclohexanone exhibits a positive Cotton effect at its carbonyl n→π* band near 290 , with the ORD curve showing an S-shaped profile and the CD displaying a positive extremum.

Applications

Stereochemistry

The Cotton effect, observed through (ORD) or (CD) measurements, plays a crucial role in determining the absolute of organic molecules containing carbonyl chromophores, particularly ketones. The octant rule provides a predictive framework for the sign of the n→π* Cotton effect in such compounds, dividing the space around the into eight octants based on a where the carbonyl carbon is at the origin, the C=O bond lies along the positive x-axis, and the plane of the carbonyl is the xy-plane. Substituents positioned in the upper-left and lower-right octants (when viewed along the x-axis with oxygen pointing away) contribute positively to the Cotton effect amplitude, while those in the lower-left and upper-right octants contribute negatively; substituents in the three nodal planes (yz, xz, and a plane bisecting the C-C(=O)-C ) have negligible influence. This empirical rule correlates the spatial arrangement of chiral centers and substituents relative to the carbonyl with the observed sign of the Cotton effect, enabling stereochemical assignments without relying on chemical methods. The octant rule has been extensively applied to assign the configurations of complex natural products, including steroids, terpenoids, and derivatives, by analyzing the sign and amplitude of their Cotton effects. In steroids, such as cholestanone derivatives, the rule confirms the at key chiral centers by comparing predicted octant contributions from ring fusions and side chains to experimental ORD curves, often resolving ambiguities in biosynthetic pathways. For terpenoids like guaiacwood alcohol, the negative Cotton effect of the derived aligns with the ()-configuration at C-4, validating the overall structure through octant . Similarly, in substituted , the rule distinguishes axial versus equatorial orientations of chiral substituents, aiding in the stereochemical elucidation of rigid ring systems. These applications underscore the rule's reliability for molecules where the carbonyl is the dominant , provided conformational rigidity minimizes octant overlap. Configurational analysis of unknown compounds frequently involves matching their ORD or curves—characterized by the , sign, and of the Cotton effect—to those of known standards with established configurations. This comparative approach is particularly effective for structurally related series, such as analogs, where similarities in spectral shape and peak positions indicate shared at critical centers, while deviations signal epimeric differences. By overlaying curves, researchers can infer the unknown's without synthesizing reference compounds, enhancing efficiency in synthetic and chemistry. An illustrative example is the use of 5α-cholestane-3β,6β-diol derivatives, such as the bis(p-dimethylaminobenzoate), in early stereochemical assignments via the exciton method, which manifests as a strong bisignate Cotton effect in CD spectra due to coupled transitions of the benzoate chromophores. The negative first Cotton effect in the 3β,6β-diol bisbenzoate confirms the anti-periplanar orientation and consistent with the 5α-series, providing a benchmark for related diols and influencing subsequent assignments in polyhydroxylated sterols. In recent years, computational methods have advanced the application of Cotton effects for determination. For instance, models trained on large datasets of computed electronic (ECD) spectra enable rapid prediction of configurations for over 10,000 chiral molecules, enhancing the efficiency of chiral analysis in and as of 2025.

Biochemistry

The Cotton effect, observed through (ORD), provides valuable insights into the secondary structures of proteins by revealing characteristic anomalous dispersion patterns associated with chiral chromophores in the backbone. For α-helical structures, a positive Cotton effect is typically observed around 222 , corresponding to the n→π* transition, which arises from the ordered helical arrangement of the polypeptide chain. In contrast, β-sheet conformations exhibit a negative Cotton effect near 230 , reflecting the extended strand and intermolecular that influence the rotational strength. These distinct ORD signatures enable the quantitative assessment of secondary structure content in proteins, often complementing (CD) data, and have been instrumental in elucidating conformational changes during folding or denaturation processes. In nucleic acids, the Cotton effect manifests in CD spectra of double-helical structures, where B-DNA displays a positive at approximately 280 nm, attributed to the cooperative base stacking interactions within the right-handed . This signal, linked to π→π* transitions of the aromatic bases, is sensitive to helical topology and environmental factors, allowing differentiation from A- or forms that show altered intensities or sign inversions. Such spectroscopic features have facilitated studies of DNA conformational dynamics and interactions with ligands or proteins. Enzyme-substrate interactions can induce extrinsic Cotton effects in ORD or CD, arising from the asymmetric perturbation of chromophores upon , which reports on the geometry and conformational shifts at . For instance, in , substrate analogs induce a Cotton effect by coordinating with the enzyme's ion, enabling mapping of the orientation relative to chiral protein elements. Similarly, to substrates or inhibitors generates ORD Cotton effects that indicate induced conformational adjustments in the enzyme's cleft. A notable application involves the β-form of silk fibroin, where ORD studies revealed a negative Cotton effect with a trough at 229–230 nm and a peak at 205 nm, confirming the antiparallel β-sheet structure in solution and distinguishing it from or helical forms. This work by Iizuka and Yang (1966) highlighted how solvent conditions modulate the conformational equilibrium, providing early evidence for β-sheet dominance in 's biomechanical properties. Recent advances as of 2025 include the application of Cotton effect analysis in studying chirogenic effects in supramolecular assemblies and pharmaceuticals, such as investigating induced Cotton effects in porphyrin-based systems for biosensing and in antidiabetic drugs like vildagliptin for conformational analysis.

Measurement

Instrumentation

Optical rotatory dispersion (ORD) measurements for the Cotton effect typically employ spectropolarimeters that scan wavelengths to capture the dispersion curve, utilizing photoelectric detection to quantify the of plane-polarized by chiral samples across a range of wavelengths. These devices incorporate a for spectral selection and a photoelectric detector, such as a , to measure the with high sensitivity. Basic digital polarimeters, like Jasco's P-2000 series, support measurements at discrete wavelengths with resolutions as fine as 0.0001° and response speeds up to 6° per second, suitable for multi-point ORD but not continuous scanning. Circular dichroism (CD) spectrophotometers generate circularly polarized light essential for detecting differential absorption by enantiomers, primarily through photoelastic modulators (PEM). The PEM, often operating at frequencies around 50 kHz, converts linearly polarized light into alternating left- and right-circularly polarized components by inducing stress-induced in a , with the resulting signal demodulated via lock-in amplification for accurate CD measurement. Commercial examples include Jasco's J-1500 series, which integrate PEM technology for robust performance in chiral analysis. Measurements of the Cotton effect occur within the ultraviolet-visible (UV-Vis) spectral range of approximately 190 to 800 , where electronic transitions in chiral molecules are prominent; this requires sample cells to ensure transparency below 220 , as absorbs in the deep UV. Solvents must also be selected for minimal absorption in this window to avoid interference. Contemporary setups frequently combine ORD and functionalities in hybrid spectropolarimeters, such as the Jasco J-1700 series (as of 2023), which adapt instrumentation for ORD via accessories like enhancements, facilitating simultaneous data collection to streamline spectral curve analysis. These integrated systems reduce setup time and improve data correlation for Cotton effect studies, with enhanced ranges down to 170 nm and up to 2600 nm for advanced applications.

Experimental Methods

Sample preparation for Cotton effect experiments requires the use of pure enantiomers dissolved in solvents that do not absorb in the region of interest to avoid with the chiral signal. Typical concentrations from 0.1 to 1 mg/mL, ensuring an optical of approximately 0.5 to 1.0 at the absorption maximum for optimal while preventing excessive light absorption. Samples must be free of impurities or aggregates, often achieved through or , and enantiomeric purity is verified prior to measurement to isolate the true Cotton effect. Scanning protocols for observing the Cotton effect involve selecting a range that encompasses the relevant electronic absorption bands of the , typically from 180 to 400 nm for UV-visible studies, with slower scan speeds (e.g., 10-50 nm/min) and integration times (e.g., 2-8 seconds per point) to capture fine spectral features. Baseline corrections are essential and performed by measuring the or alone under identical conditions, then subtracting this from the sample data to eliminate instrumental and contributions. Multiple accumulations (3-10 scans) are often averaged to enhance signal quality, particularly near absorption edges where noise is higher. Avoiding artifacts is critical for reliable Cotton effect data, as birefringence in sample cells can introduce false rotational signals, which is mitigated by using strain-free quartz cuvettes and verifying flat baselines with solvent scans. Stray light, which can distort measurements at high absorbances, is minimized by maintaining low optical densities and ensuring proper instrument alignment, while temperature control (typically at 20-25°C with ±0.1°C stability) ensures reproducibility by preventing thermal fluctuations in molecular conformation. Nitrogen purging of the optical path is recommended for far-UV measurements to reduce oxygen absorption artifacts. Data interpretation of Cotton effects in complex molecules often requires deconvolution techniques to resolve overlapping bands from multiple chromophores, using methods like Gaussian or to extract individual contributions. For instance, in proteins or natural products, software-based separates secondary structure signals in far-UV CD spectra, allowing assignment of positive or negative Cotton effects to specific transitions. These approaches prioritize regions of minimal overlap and validate results against known spectra for accuracy.

References

  1. [1]
    Cotton Effect - an overview | ScienceDirect Topics
    Cotton effect (CE) refers to the anomalous optical rotatory dispersion (ORD) observed in optically active compounds, characterized by the presence of a ...Missing: history | Show results with:history
  2. [2]
  3. [3]
    Chiroptical Sensing: A Conceptual Introduction - PMC
    Chiroptical responses have been an essential tool over the last decades for chemical structural elucidation due to their exceptional sensitivity to geometry ...
  4. [4]
    Basic Principles and Relevant Applications in Chirality Sensing - MDPI
    The traditional method of structural research is the empirical rules-based correlation between the sign and value of the Cotton effect observed at a given ...
  5. [5]
    Circular Dichroism - an overview | ScienceDirect Topics
    The major advantages of CD are it is fast (minutes compared to hours), it is non-destructive, and it uses relatively inexpensive equipment. Calculation of ...
  6. [6]
    What Are the Advantages and Limitations of Circular Dichroism (CD ...
    Advantages of Circular Dichroism Spectroscopy · 1. Rapid Structural Analysis with Minimal Sample Consumption · 2. Real-Time Monitoring of Conformational Changes.
  7. [7]
    Circular Dichroism Spectroscopy - an overview | ScienceDirect Topics
    β-sheet shows a strong positive band at 195–200 nm, and a broad negative peak at 216–218 nm indicates the presence of β-sheet. There is a strong negative band ...
  8. [8]
    Circular dichroism in functional quality evaluation of medicines
    Circular dichroism (CD) is a non-destructive and powerful technique for providing structure and ligand interaction information of small molecules as well as ...<|control11|><|separator|>
  9. [9]
    Determination of Absolute Configuration Using Chiroptical Methods
    Oct 2, 2013 · This chapter is aimed at covering the current state-of-the-art of chiroptical spectroscopy in combination with quantum chemical calculations to determine the ...
  10. [10]
    The First Decades after the Discovery of CD and ORD by Aimé ...
    Feb 17, 2012 · Oddly, this author maintains that “Aimé Cotton (1872–1944) discovered the [Cotton] effect while Professor of Physics at the Sorbonne in Paris.” ...
  11. [11]
    The First Decades after the Discovery of CD and ORD by Aimé ...
    This scenario is similar to the so-called Cotton effect, which was first described in 1895. The tandem choice of both the CP light sense and its wavelength ...Missing: original | Show results with:original
  12. [12]
    [PDF] Chiroptical spectroscopy - Merten Lab
    Aimé Cotton. (1869-1951). Page 12. Chiroptical spectroscopy | Prof. Dr. C ... The Cotton effect. Change of OR near absorbance band positive Cotton effect ...
  13. [13]
    [PDF] PART I - Wiley Monthly Title Update and Image Download Site
    The story of the Cotton effect begins with its discovery in 1895. Although the news was hailed by leading physicists and chemists, studies to extend, exploit, ...
  14. [14]
    When Stereochemistry Raised Its Ugly Head in Coordination ... - MDPI
    Sep 12, 2020 · This is called the Cotton effect and refers to the change in sign of the optical rotatory dispersion close to an absorption band. Close to the ...
  15. [15]
    Optical Activity in Coordination Chemistry - ACS Publications
    Nov 4, 1994 · The demonstration of the optical activity of octahedral complexes was important in confirming Alfred Werner's intuitive ideas about ...Missing: compounds Cotton
  16. [16]
    Thomas Martin Lowry, 1874-1936 - Journals
    In recent years, Lowry extended his work on rotatory dispersion from transparent to absorbing media, i.e., to a study of the Cotton effect, with considerable ...
  17. [17]
    Optical Rotatory Dispersion - an overview | ScienceDirect Topics
    Its application to organic molecules awaited the pioneering work of Mitchell, Lowry, and Kuhn in the late 1920s and 1930s. The development of CD instrumentation ...Missing: advancements | Show results with:advancements
  18. [18]
    Optical Rotatory Dispersion - an overview | ScienceDirect Topics
    When measured inside of an absorption band, the ORD will exhibit anomalous dispersion, which is referred to as a Cotton effect. A positive Cotton effect ...
  19. [19]
    19.9: Optical Rotatory Dispersion and Circular Dichroism
    Mar 5, 2021 · In fact, when a substance exhibits a Cotton effect, not only does it transmit clockwise and counterclockwise circularly polarized light with ...Missing: history | Show results with:history
  20. [20]
    [PDF] OPTICAL ROTATORY DISPERSION (ORD)
    • Circular Dichroism. • Instrumentation. • Cotton Effect. • Applications of ORD ... ❖The Cotton effect is the characteristic change in optical rotatory ...Missing: chiroptical | Show results with:chiroptical
  21. [21]
    Beginners guide to circular dichroism | The Biochemist
    Mar 26, 2021 · The CD significance of this is that although the best matching proteins have βII structure, the spectrum of βIIs look like random coil spectra ...Challenges To Acquiring A... · Interpreting A Circular... · Some Recent Applications Of...
  22. [22]
    Circular Dichroism Spectroscopy - JASCO
    Circular Dichroism spectroscopy exploits the fundamental property described by the 'Cotton Effect'. See here the Wikipedia article. Circularly polarized ...
  23. [23]
    [PDF] Circular Dichroism
    Sep 20, 2004 · CD spectra are coincident with the absorption spectra of molecules. However, ORD spectra are dispersive and have long "tails" of rotation to.
  24. [24]
    Optical Rotatory Dispersion - an overview | ScienceDirect Topics
    2 The Cotton effect circular dichroism and optical rotatory dispersion curves. An ... The strength of the electric dipole is related to the electric dipole ...
  25. [25]
    Chirality and Chiroptical Effects in Metal Nanostructures ...
    Aug 1, 2017 · ... optical rotation changes sign. This particular sign change is known as the Cotton effect.74, 75 It originates from the Kramers-Kronig ...
  26. [26]
    Kramers−Kronig Transformation for Optical Rotatory Dispersion Studies
    ### Summary of Kramers-Kronig Formula from https://pubs.acs.org/doi/10.1021/jp0524328
  27. [27]
    [PDF] Application Note #11 - BioLogic
    ORD (Optical Rotatory Dispersion) accessory using the MOS450 AF/CD spectrometer ... This anomalous rotation is called the Cotton Effect [3]. Indirectly ...
  28. [28]
    [PDF] 6 Chirooptical Methods - Thieme Connect
    In the region of absorption bands the normal ORD curve is overlaid with an S-shaped component to produce the so-called anomalous ORD curve (Fig. 1.39). |E1| |Er ...
  29. [29]
    Principles of CD/ORD (3): Features of ORD, CD measurements
    Oct 6, 2020 · Cotton effect Fig. 6 schematically shows the relationship between the CD/ORD spectra and the UV spectrum in the absorption wavelength region ...<|separator|>
  30. [30]
    Octant rule: Definition, application with examples - Chemistry Notes
    Nov 9, 2023 · ... positive cotton effect as it lies in the ... Plane A passes through the carbonyl group and carbon number 4 bisecting the cyclohexanone ...
  31. [31]
    The octant rule. XX : Synthesis and circular dichroism of (1s,5s ...
    The octant rule. XX : Synthesis and circular dichroism of (1 s ,5 s )-dimethyladamantan-2-one - predicted to have zero cotton effect · Abstract · References (23).Missing: seminal paper
  32. [32]
    New Cotton Effects in Polypeptides and Proteins - ACS Publications
    Substrate- and Inhibitor-induced Changes in the Optical Rotatory ... Extrinsic Cotton Effects and the Mechanism of Enzyme Action. 1965, 37-104 ...<|control11|><|separator|>
  33. [33]
    Conformational studies on the interaction of chymotrypsin with ...
    The individual constants, the shape of the plots and the Cotton effects all indicate that conformational changes are induced by substrates and inhibitors of ...
  34. [34]
    P-2000 High Accuracy Digital Polarimeter - JASCO Inc
    The P-2000 is a configurable, high-precision polarimeter with a 6°/s speed, 0.0001° resolution, and a wide dynamic range of 0.001° to ±90°.
  35. [35]
    CD Instrumentation FAQs - Applied Photophysics
    What is a photoelastic modulator (PEM)?. The photoeleastic modulator (PEM) converts horizontal linearly polarised light to circularly polarized light. The PEM ...
  36. [36]
    Electronic Circular Dichroism - Encyclopedia.pub
    Circular dichroism (CD) is defined as the differential absorption of left- and right-circularly polarized electromagnetic radiation by a sample.<|control11|><|separator|>
  37. [37]
    [PDF] J-810 Spectropolarimeter - University of Manitoba
    The J-810 is a CD spectropolarimeter, a hybrid instrument measuring circular dichroism (CD) and optical rotatory dispersion (ORD) with low straylight.
  38. [38]
    [PDF] Sample Preparation for Circular Dichroism Measurements
    When preparing a sample for CD measurements the absorption of light must be consid ered. For normal light the optical density (OD) of the sample is given by ...Missing: experimental methods Cotton effect ORD
  39. [39]
    [PDF] Sample Preparation for CD Measurements
    Volume requirements: For the 1 mm path-length cuvette, 400μL is required to ensure beam coverage. Spin down or filter samples to remove particulates. Cuvettes:.Missing: methods Cotton effect ORD
  40. [40]
    [PDF] Circular Dichroism - Jasco Spain
    at 210 nm. The Z-form structure is a left-handed double helix that displays a negative band at 290 nm, a positive peak at 260 nm, and negative maximum ~200 nm.
  41. [41]
    Mastering Circular Dichroism: Techniques, Troubleshooting ...
    Baseline Correction: Subtract the background signal from the raw CD data to obtain the true spectral curve of the protein.
  42. [42]
    Using circular dichroism spectra to estimate protein secondary ... - NIH
    For example, α-helical proteins have negative bands at 222 nm and 208 nm and a positive band at 193 nm. ... The peptide references sets were 1, α-helix, β- ...
  43. [43]
    SESCA: Predicting Circular Dichroism Spectra from Protein ...
    Aug 11, 2019 · We introduce a new computational method to calculate the electronic circular dichroism spectra of proteins from a structural model or ensemble.<|separator|>
  44. [44]