User:Brownaj77

From Wikipedia, the free encyclopedia

Magnetically induced Circular Dichroism (MCD) is the differential absorption of left and right circularly polarized light, induced in a sample by a strong magnetic field oriented parallel to the direction of light propagation. In general, this means wrapping a circular dichrometer in a large electromagnet. If the magnet is water cooled, field strength of a couple of teslas are common. With a superconducting magnet, field strengths can reach 7 Teslas.

It was first shown by Faraday that optical activity could be induced in matter by a longitudinal magnetic field (a field in the direction of light propagation) [1]. The development of MCD really began in the 1930s when a quantum mechanical theory of MOR (magnetic optical rotatory dispersion) in regions outside absorption bands was formulated. The expansion of the theory to include MCD and MOR effects in the region of absorptions, which were referred to as “anomalous dispersions” was developed soon thereafter. There was, however, little effort made to refine MCD as a modern spectroscopic technique until the early 1960s. Since that time there have been numerous studies of MCD spectra for a very large variety of samples, including stable molecules in solutions, in isotropic solids, and in the gas phase, as well as unstable molecules entrapped in noble gas matrices. More recently, MCD has found useful application in the study of biologically important systems including metalloenzymes and proteins containing metal centers.[2] [3]

The main focus of the MCD is to analyze the A, B, C terms of the spectra. Principles and examples will be explained later on this page.

Applicable Cases[edit]

MCD can be used as an optical technique for the detection of electronic structure of both the ground states and excited states. It is also a strong addition to the more commonly used absorption spectroscopy, and there are two reasons that explain this. First, a transition buried under a stronger transition can appear in MCD if the first derivative of the absorption is much larger for the weaker transition or it is of the opposite sign. Second, MCD will be found where no absorption is detected at all if ΔA > (ΔAmin) but A<Amin, where (ΔA)min and Amin are the minimum of ΔA and A that are detectable. Typically, (ΔAmin) and Amin are of the magnitudes around 10-5 and 10-3 respectively. So, a transition can only be detected in MCD, not in the absorption spectroscopy, if ΔA/A > 10-2. This happens in paramagnetic systems that are at lower temperature or that have sharp lines in the spectroscopy.[4]

Differences between CD and MCD[edit]

In natural optical activity, the difference between the lcp light and the rcp light is caused by the dissymmetry of the molecules. Due to the handedness of the molecule, the absorption of the lcp light would be different from the rcp light. However, in MCD in the presence of a magnetic field, lcp and rcp no longer interact equivalently with the absorbing medium. Thus there is not the same direct relation between magnetic optical activity and molecular stereochemistry which would be expected, because it is found in natural optical activity. So, natural CD is a relatively rare phenomenon, while MCD is a property of all matter.[5]

Spectrometer[edit]

Modern day spectrometers can achieve simultaneous measurement of absorbance and the MCD ΔA along the same light path.[6] This eliminates error introduced through multiple measurements or different instruments that previously occurred before this advent. A MCD spectrometer begins with a light source that emits a monochromatic wave of light. This wave is passed through a Rochon prism linear polarizer, which separates the incident wave into two beams that are linearly polarized by 90 degrees. The two beams split into different paths- one beam (the extraordinary beam) traveling directly to a photomultiplier (PMT), and the other beam (the ordinary beam) passing through a photoelastic modulator (PEM). The PMT for the extraordinary beam detects the light intensity of the input beam, while the PEM causes a 1/4 wavelength shift that converts linearly polarized light into circularly polarized light. Linearly polarized light has two circular components with intensity represented by:



The PEM will delay one component of linearly polarized light with a time dependence that advances the other component by 1/4 λ (hence, quarter-wave shift). The departing circularly polarized light oscillates between rcp and lcp in a sinusoidal time-dependence as depicted below:



The light finally travels through a magnet containing the sample, and the transmittance is recorded by another PMT. The schematic is given below:



The intensity of light from the ordinary wave that reaches the PMT is governed by the equation:



A- and A+ are the absorbances of lcp or rcp, respectively; ω is the modulator frequency- usually a high acoustic frequency such as 50 kHz; t is time; and δ0 is the time dependent wavelength shift.

This intensity of light passing through the sample is converted into a two component voltage via a current/voltage amplifier. A DC voltage will emerge corresponding to the intensity of light passed through the sample. If there is a ΔA, then a small AC voltage with be present that corresponds to the modulation frequency, ω. From such voltage, ΔA and A can be derived using the following relations:


File:The measument of det A.png


File:The measument of A.png


where Vex is the voltage measured by the PMT from the extraordinary wave.



Photograph given by permission of a low-temperature MCD Spectrometer built at the University of Michigan.

MCD signals tend to be small, so the measurements often take time. To get a reasonable signal to noise ratio, the experiments can often take a few hours. Multiple spectra are often taken, digitized, accumulated and stored via computer. Although there is much overlap in the requirements and use of instruments, ordinary CD instruments are usually optimized for operation in the ultraviolet, approximately 170–300 nm, while MCD instruments are typically required to operate in the visible to near infrared, approximately 300–2000 nm.

The physical processes that lead to MCD are substantively different from those of CD. However, like CD, it is dependent on the differential absorption of left and right hand circularly polarized light. MCD will only exist at a given wavelength if the molecule has an optical absorption at that wavelength. This is distinctly different from the related phenomenon of Optical Rotatory Dispersion (ORD), which can be observed at wavelengths far from any absorption band.

Some superconducting magnets have a small sample chamber, far too small to contain the entire optical system. Instead, the magnet sample chamber has windows on two opposite sides. Light from the source enters one side, interacts with the sample (usually also temperature controlled) in the magnetic field, and exits through the opposite window to the detector. Optical relay systems that allow the source and detector each to be about a meter from the sample are typically employed. This arrangement avoids many of the difficulties that would be encountered if the optical apparatus had to operate in the high magnetic field, and also allows for a much less expensive magnet.

How it works[edit]

Basically, we consider the systems are of localized, non-interacting absorbing centers. Based on the semi-classical radiation absorption theory within the electric dipole approximation, the electric vector of the Circular Polarized waves propagates in the + z direction, where + and -denote to rcp light and lcp light respectively.

File:The formula description of the circular light.png

In this system, ω = 2nν is the circular frequency, and Ii = n-ik is the complex refractive index. As the light travels, the attenuation of the beam is expressed by



in which k means the absorption coefficient of the medium in the z direction Circular Dichroism (CD) is then defined through


File:Different absorpsion in terms of k.png


And it follows the sign convention of natural optical activity. In the presence of the static, uniform external magnetic field, The Hamiltonian for the absorbing center is of the form


File:Hamiltonian of the state in MCD.png


which is parallel to the direction of the propagation of the electric field k.[7] Due to the transition between the two eigenstates of H0, a and j, the absorption and the CD intensity would be determined as follows:


File:Intensity in terms of k.png

Discrete line spectrum[edit]

In a certain frequency range, if the transitions of the absorbing centers are relatively sparse, the absorption spectrum in this region would be of a series of lines. Each line would correspond to a single transition and be either completely or partially resolved from neighboring lines. Since it is of the discrete form, then the absorption coefficient and CD of the line are given by

When the Zeeman Effect is small compared to zero-field state separations, line width, and kT and when the line shape is independent of H, it could be obtain to first order in H for the expression of Δk, assuming J to be sufficiently high in energy that Nj = 0. And it is given by:

Δk is the field-dependent difference between lcp and rcp absorption, α is the electric permeability, n is the index of refraction, and H is the applied magnetic field, k is the Boltzmann constant and T is the temperature. Further, f is a band shape function and f’ is the correspondent first derivative. D is the dipole strength parameter, and it is defined as[8]

File:Definition of D term.png

There are various form of the formulas above. A more common form is in terms of A (absorption), which would be:

The constants A, B, and C are characteristic parameters specific to a given molecule and to a particular transition, which are referred to as A, B, and C terms, respectively, a nomenclature introduced by Stephens and dating back to Serber's early theory of MOR.


A term arises due to Zeeman splitting of the ground or excited degenerate states, like what is shown in the (2) in Figure 1 The small Zeeman splitting would make the oppositely signed transitions rcp light and lcp light almost cancel out with each other, which leads to the derivative shape of the band. B term is due to the field-induced mixing of states. As shown in (3) of the Figure 1, in order for there to be an appearance of B term, there must be another state K, which is closely related in energy to either the ground state or the excited states. Also, the energy between the ground state and the excited state should be high enough so that the excited state is not highly populated. Usually, B term would have the absorption band shape. Otherwise, it will lead to the C term (4) of Figure 1, which requires the degeneracy of the ground state. This happens due to a change in the population of molecules over the Zeeman sublevels of a paramagnetic ground state. In addition, the C term is observed only for molecules with ground state paramagnetism. Importantly, A and B terms are independent of temperature while the C term is dependent on temperature. [9]Decrease of the temperature and increase of the magnetic field would further increase the C term intensity until it reaches the maximum(saturation limit). Experimentally, the C term spectrum can be obtained from MCD raw data by subtraction of MCD spectra measured in the same applied magnetic field at different temperatures, while A and B terms can be distinguished via their different band shapes. [10] The relative contributions of A, B and C terms to the MCD spectrum are proportional to the inverse line width, energy splitting, and temperature:

in which, ΔΓ is line width and ΔE is zero-field state separation magnitudes. Typically, the values (ΔГ= 1000 cm-1, ΔE = 10,000 cm-1 , and kT = 6 cm-1 (at 10 K), the three terms will make relative contributions 1:0.1:150. So, at low temperature the C term will dominate over A and B term.[11]

Example on C terms[edit]

In the visible and near-ultraviolet regions, the hexacyanoferrate(III) ion exhibits three strong transitions, which have been ascribed to Ligand to Metal Charge Transfer (LMCT) transitions. The three intense bands are located at 24500, 32700, and 40500 cm-1 in the spectra of the Fe (III) complex Fe(CN)63-, which are all found lower in energy than the lowest energy intense band for the Fe(II) complex Fe(CN)62- found at 46000 cm-1.[12] The red shift as the oxidation state of the metal increases is characteristic of LMCT bands.


The dependence of the C term of MCD is explored here. The ground state of the anion is 2T2g, which derives from the electronic configuration (t2g)5. So, there would be an unpaired electron in the d orbital of Fe3+ From that, we can assign the three bands correspondent to the transitions 2t2g2t1u1, 2t2g2t1u2, 2t2g2t2u. Noticed that two of the excited states are of the same symmetry, and based on the group theory, we can assert that they would actually mix with each other so that there are no pure σ, π character of the two t1u states, but for t2u, there would be no inter-mixing. It should be noted that A terms are also possible from the degenerate excited states, but the temperature-dependence studies showed that they are not as dependent as the C term.[13]


A more recent study was done on the MCD of Fe(CN)63- embedded in a thin poly(vinyl alcohol) (PVA) film. We can clearly see that there is a dependence on temperature by the C term. With the equations that were given in the theory part, we can get the appropriate expression for C0 and D0. The observed values of C0/D0 at room temperature for the three bands in the Fe(CN)63- spectrum are 1.2, -0.6, and 0.6 (with some experimental uncertainty in the last two values) respectively. Thus the signs we got from the experiment for the three C terms in the MCD spectra (positive, negative, and positive) establish the energy ordering as: 2t2g2t1u2<2t2g2t2u<2t2g2t1u1

Example on A and B terms[edit]

To have an A- and B-term in the MCD spectrum, a molecule must contain degenerate excited states (A-term) and excited states close enough in energy to allow mixing (B-term). One case exemplifying these conditions is a square planar, d8 complex such as [(n-C4H9)4N]2Pt(CN)4.[14] In addition to containing A- and B-terms, this example demonstrates the effects of spin-orbit coupling in Metal to Ligand Charge Transfer transitions. As shown in Figure 1, the molecular orbital diagram of [(n-C4H9)4N]2Pt(CN)4 reveals MLCT into the antibonding π* orbitals of cyanide. The ground state is diamagnetic (thereby eliminating any C-terms) and the LUMO is the a2u. The dipole allowed MLCT transitions are the a1g  a2u and the eg  a2u. Another transition, b2u  a2u, is a weak and orbitally forbidden singlet, but still can be observed in MCD. [14]


Because A- and B-terms arise from the properties of states, all of the singlet and triplet excited states are given in Figure 2.


Mixing of all these singlet and triplet states will occur and is attributed to the spin orbit coupling of platinum’s 5d orbitals (ζ ~ 3500cm-1), as shown in figure 3. The black lines on the figure indicate the mixing of 1A2u with 3Eu to give two A2u states. The red lines show the 1Eu, 3Eu, 3A2u, and 3B1u states mixing to give four Eu states. The blue lines indicate remnant orbitals after spin-orbit coupling that are not a result of mixing.


At this point in the example, it’s prudent to state that the assignment of the [(n-C4H9)4N]2Pt(CN)4 MCD spectrum has been debatable. The mixing of orbitals via spin-orbit coupling shown in Figure 3 was determined by diagonalization of spin-orbit secular determinants. However, due to the close agreement of these calculated results with the experimental spectrum, the hypothesis of spin-orbit coupling is given credence. The labeled experimental spectrum is shown in Figure 4.


The A-terms are derivative in shape, while the B-terms appear as broad absorption peaks. A molecular state depiction of the transitions seen in the spectrum is given by Figure 5.


The A-term assigned at 35.400kcm-1 is the transitions from ground A1g to the first Eu state. Another positive A-term arises at 38.400kcm-1 and is the transition from A1g to the second Eu state. The mixing of the second Eu state with the A2u state yields overlapping B-terms that are opposite in sign and present what is known as a pseudo A-term. The same situation happens again at 45.500kcm-1. A positive A-term emerges for the transition between the a1g¬ to the fourth highest Eu state on the MO digram (Figure 3). Another pseudo A-term comes from B-terms of Eu and A2u. The two B1u states are assigned to the shoulder bands at ~ 42.000kcm-1 and ~47.000kcm-1. This example for the A- and B-terms shows a very real scenario in which the interpretation and assignment of the spectrum is not straight forward. Platinum’s spin orbit coupling mixes the MLCT excited states and necessitates the use of calculation. However, this example does show that MCD can be used in a complex system as a viable method to extrapolate the relative ordering of d-orbitals. From this data pertaining to [(n-C4H9)4N]2Pt(CN)4, the relative order of the d-orbitals were determined in decreasing energy as: b1g (dx2- y2) > a1g (dz2) > eg (dxy,dyz) > b2g (dxy).

Other fields[edit]

In biology, metalloproteins are the most likely candidates for MCD measurements, as the presence of metals with degenerate energy levels leads to strong MCD signals. In the case of ferric heme proteins[15], MCD is capable of determining both oxidation and spin state to a remarkably exquisite degree. In regular proteins, MCD is capable of stoichiometrically measuring the tryptophan content of proteins, assuming there are no other competing absorbers in the spectroscopic system. In addition,the application of MCD spectroscopy greatly improved the level of understanding in the ferrous non-heme systems because of the direct observation of the d~d transitions, which generally can not be obtained in optical absorption spectroscopy owing to the weak extinction coefficients and are often electron paramagnetic resonance silent due to relatively large ground-state sublevel splittings and fast relaxation times.[16]

References[edit]

  1. ^ A. D. Buckingham and P. J. Stephens, Ann. Rev. Phys. Chem. 1966 17,399
  2. ^ W. Roy Mason, Practical Guide to Magnetic Circular Dichroism Spectroscopy, A, 2007
  3. ^ P. N. Schatz, A. J. McCaffery The Faraday Effect
  4. ^ Stephens,P. J. Annu. Rev. Phys. Chem. 1974, 25, 201.
  5. ^ Dennis Caldwell, Annu. Rev. Phys. Chem. 1971. 22, 259-278
  6. ^ G. A. Osborne, The Review of Scientific Instruments, 1973, 44,
  7. ^ Buckingham,Ann. Rev. Phys. Chem. 17, 399
  8. ^ P. J. Stephens, Annu. Rev. Phys. Chem. 1974.25, 201-232
  9. ^ Lehnert, N.; DeBeer George, S.; Solomon, E. I. Curr. Opin. Chem. Biol. 2001, 5, 176-187.
  10. ^ Solomon, E. I.; Pavel, E. G.; Loeb, K. E.; Campochiaro, C. Coord. Chem. Rev. 1995, 144, 369-460
  11. ^ Neese, F.; Solomon, E. I. Inorg. Chem. 1999, 38, 1847-1865.
  12. ^ Stephens, P. J. Inorg. Chem. 1965,4, 1690-1692
  13. ^ Upton,A. H. P.; Williamson, B. E., J. Phys. Chem., 1994, 98, 71
  14. ^ Isci, H.; Mason, W. R. Inorg. Chem. 1975, 14, 905
  15. ^ Giorgio Zoppellaro, Biopolymers 2009, 91, 1064
  16. ^ E.I. Solomon et al.Coordination Chemistry Reviews,144 1995 369-460

See also[edit]

Category:Polarization Category:Spectroscopy