Resonance Raman spectroscopy

Last updated
Energy level diagram showing relationship between Rayleigh, Raman, and resonance Raman scattering and fluorescence. Energy levels in Raman spectroscopy.jpg
Energy level diagram showing relationship between Rayleigh, Raman, and resonance Raman scattering and fluorescence.

Resonance Raman spectroscopy (RR spectroscopy or RRS) is a variant of Raman spectroscopy in which the incident photon energy is close in energy to an electronic transition of a compound or material under examination. [1] This similarity in energy (resonance) leads to greatly increased intensity of the Raman scattering of certain vibrational modes, compared to ordinary Raman spectroscopy.

Contents

Resonance Raman spectroscopy has much greater sensitivity than non-resonance Raman spectroscopy, allowing for the analysis of compounds with inherently weak Raman scattering intensities, or at very low concentrations. [2] [3] It also selectively enhances only certain molecular vibrations (those of the chemical group undergoing the electronic transition), which simplifies spectra. [3] For large molecules such as proteins, this selectivity helps to identify vibrational modes of specific parts of the molecule or protein, such as the heme unit within myoglobin. [4] Resonance Raman spectroscopy has been used in the characterization of inorganic compounds and complexes, [5] proteins, [6] [7] nucleic acids, [8] pigments, [8] and in archaeology and art history. [8]

Theory

In Raman scattering, photons collide with a sample and are scattered with a difference in energy: The scattered photons may be higher or lower in energy (have a shorter or longer wavelength) than the incident photons. This difference in energy is caused by excitation of the sample to a higher or lower vibrational energy level: if the sample was initially in an excited vibrational state, the scattered photon may be higher in energy than the incident photon (anti-Stokes Raman scattering). Otherwise, the scattered photon has a lower module of energy than the incoming photon (Stokes Raman scattering). Among the two phenomena, Stokes shift and anti-Stokes shift, the former is the most likely to occur. As a consequence, the relative intensity of Raman spectra acquired in Stokes mode is more intense than the other. For most materials, Raman scattering is extremely weak compared to Rayleigh scattering, in which light is scattered without loss of energy. [9] Raman-scattered light, which contains information about vibrational transitions, is therefore difficult to observe for many substances.

Resonance Raman spectroscopy takes advantage of an increase in the intensity of Raman scattering when the incident photons match the energy of an electronic transition. If the energy of the photon striking the sample is equal or close to that of an electronic transition in the sample, certain Raman-active vibrational modesthose producing nuclear displacement in the same direction as the electronic transition [10] will exhibit greatly enhanced scattering, up to 106-fold compared to nonresonance Raman. [3] For totally symmetric modes, this increased scattering intensity results from so-called A-term or Franck-Condon scattering, due to the nonzero Franck-Condon overlaps between ground and excited states. Nontotally symmetric modes may also be enhanced by B-term or Herzberg-Teller scattering, if the symmetry of the mode is contained in the direct product of the two electronic state symmetries. [11] Resonance enhancement is most apparent in the case of π-π* transitions and least for metal centered (d–d) transitions. [5] Like ordinary Raman spectroscopy, RRS observes vibrational transitions producing a nonzero change in the polarizability of the molecule or material being studied.

Resonance Raman scattering differs from fluorescence in that it occurs without vibrational relaxation during the lifetime of the excited electronic state. It thus exhibits much narrower line widths than fluorescence. [11] However, fluorescence and resonance Raman scattering co-occur in many materials, and interference from fluorescence may complicate the collection of resonance Raman spectra. [3]

Variants

Typically, resonance Raman spectroscopy is performed in the same manner as ordinary Raman spectroscopy, using a single laser light source to excite the sample. The difference is the choice of the laser wavelength, which must be selected to match the energy of an electronic transition in the sample. A tunable laser is thus often used for resonance Raman spectroscopy, since a single laser can be used to generate many possible excitation wavelengths to match different samples. [8] By using multiple lasers, pulsed lasers, and/or certain sample preparation techniques, a range of more sophisticated variants of RRS can be performed, including:

Applications

Resonance (top) and nonresonance (bottom) Raman spectra of MoS2 on silicon. Note that excitation at 633 nm, near an electronic transition, causes appearance of bands that are too faint to be visible with excitation at 532 nm. Figure courtesy of David Tuschel. MSWFig6.webp
Resonance (top) and nonresonance (bottom) Raman spectra of MoS2 on silicon. Note that excitation at 633 nm, near an electronic transition, causes appearance of bands that are too faint to be visible with excitation at 532 nm. Figure courtesy of David Tuschel.

Because of its selectivity and sensitivity, resonance Raman spectroscopy is typically used to study molecular vibrations in compounds that would have very weak and/or complex Raman spectra in the absence of resonance enhancement. Like ordinary Raman spectroscopy, resonance Raman is compatible with samples in water, which has a very weak scattering intensity and little contribution to spectra. However, the need for an excitation laser with a wavelength matching that of an electronic transition in the analyte of interest somewhat limits the applicability of the method. [8]

Pigments and Dyes

Dyes and pigments, all of which exhibit electronic transitions in the visible part of the electromagnetic spectrum, were among the first substances to be studied by resonance Raman spectroscopy. Resonance Raman spectra of beta-carotene and lycopene in intact plant samples were reported in 1970. [8] Since then, the method has been used to noninvasively measure levels of these nutrients in human skin. [19] The resonance Raman spectra of other polyene pigments, such as spheroidene and retinal, have been used to identify differences in chromophore conformation in photoactive proteins. [20] [21] Resonance Raman spectroscopy has been used in archaeology to identify dyes and pigments in cultural artifacts, and the ability of RRS to distinguish different modern inks and dyes has found application in forensic science. [8]

Proteins

Proteins have been widely examined by resonance Raman spectroscopy. Protein-bound cofactors that absorb in the visible wavelength range, such as heme, flavins, or transition metal complexes, can be examined by RRS with minimal spectral overlap from the rest of the molecule. [7] [22] This method has been used to examine gas binding in hemeproteins [23] and the catalytic cycle of various enzymes. [24] Using ultraviolet laser excitation, it is possible to selectively excite the sidechains of aromatic amino acids (phenylalanine, tyrosine, and tryptophan) to deduce the local environment and hydrogen-bonding interactions by these residues. [25] With shorter-wavelength ("deep") ultraviolet excitation, it is also possible to excite the peptide bonds of a protein in order to examine secondary structure. Protein folding and denaturation have been examined using deep-UV resonance Raman spectroscopy of the polypeptide backbone, with excitation wavelengths shorter than 200 nm. [25]

Nucleic acids and viruses

Resonance Raman spectroscopy with ultraviolet excitation can be used to examine the chemistry, structure, and intermolecular interactions of nucleic acids, specifically the bases. Interactions between nucleic acids and DNA-binding compounds such as drugs can be examined by selectively exciting either the nucleobases or the drug itself. [8] The resonance Raman spectra of DNA can be used to identify bacterial DNA in living cells, and to quantitate DNA under different culture conditions, and even to distinguish different bacterial species. [8] Viruses have also been studied using UV resonance Raman spectroscopy; the method has the capability to separately interrogate the structure of the nucleic acid or capsid protein components of the virus, through the choice of the appropriate excitation wavelength. [26]

Nanomaterials

Resonance Raman spectroscopy has also been used to characterize the structure and photophysical properties of nanoparticles. Using lasers tuned to the visible and near-infrared electronic transitions of carbon nanotubes, it is possible to enhance structure-sensitive vibrational bands of the nanotubes. [8] Nanowires of inorganic semiconductor materials including gallium phosphide and carbon-encapsulated mercury telluride have also been shown to exhibit resonance Raman spectra with visible excitation light. [27] [28]

See also

Related Research Articles

<span class="mw-page-title-main">Infrared spectroscopy</span> Measurement of infrared radiations interaction with matter

Infrared spectroscopy is the measurement of the interaction of infrared radiation with matter by absorption, emission, or reflection. It is used to study and identify chemical substances or functional groups in solid, liquid, or gaseous forms. It can be used to characterize new materials or identify and verify known and unknown samples. The method or technique of infrared spectroscopy is conducted with an instrument called an infrared spectrometer which produces an infrared spectrum. An IR spectrum can be visualized in a graph of infrared light absorbance on the vertical axis vs. frequency, wavenumber or wavelength on the horizontal axis. Typical units of wavenumber used in IR spectra are reciprocal centimeters, with the symbol cm−1. Units of IR wavelength are commonly given in micrometers, symbol μm, which are related to the wavenumber in a reciprocal way. A common laboratory instrument that uses this technique is a Fourier transform infrared (FTIR) spectrometer. Two-dimensional IR is also possible as discussed below.

<span class="mw-page-title-main">Spectroscopy</span> Study involving matter and electromagnetic radiation

Spectroscopy is the field of study that measures and interprets the electromagnetic spectra that result from the interaction between electromagnetic radiation and matter as a function of the wavelength or frequency of the radiation. In simpler terms, spectroscopy is the precise study of color as generalized from visible light to all bands of the electromagnetic spectrum.

<span class="mw-page-title-main">Raman spectroscopy</span> Spectroscopic technique

Raman spectroscopy is a spectroscopic technique typically used to determine vibrational modes of molecules, although rotational and other low-frequency modes of systems may also be observed. Raman spectroscopy is commonly used in chemistry to provide a structural fingerprint by which molecules can be identified.

The term biophotonics denotes a combination of biology and photonics, with photonics being the science and technology of generation, manipulation, and detection of photons, quantum units of light. Photonics is related to electronics and photons. Photons play a central role in information technologies, such as fiber optics, the way electrons do in electronics.

<span class="mw-page-title-main">Fluorescence spectroscopy</span> Type of electromagnetic spectroscopy

Fluorescence spectroscopy is a type of electromagnetic spectroscopy that analyzes fluorescence from a sample. It involves using a beam of light, usually ultraviolet light, that excites the electrons in molecules of certain compounds and causes them to emit light; typically, but not necessarily, visible light. A complementary technique is absorption spectroscopy. In the special case of single molecule fluorescence spectroscopy, intensity fluctuations from the emitted light are measured from either single fluorophores, or pairs of fluorophores.

In physics and physical chemistry, time-resolved spectroscopy is the study of dynamic processes in materials or chemical compounds by means of spectroscopic techniques. Most often, processes are studied after the illumination of a material occurs, but in principle, the technique can be applied to any process that leads to a change in properties of a material. With the help of pulsed lasers, it is possible to study processes that occur on time scales as short as 10−16 seconds. All time-resolved spectra are suitable to be analyzed using the two-dimensional correlation method for a correlation map between the peaks.

<span class="mw-page-title-main">Raman scattering</span> Inelastic scattering of photons

Raman scattering or the Raman effect is the inelastic scattering of photons by matter, meaning that there is both an exchange of energy and a change in the light's direction. Typically this effect involves vibrational energy being gained by a molecule as incident photons from a visible laser are shifted to lower energy. This is called normal Stokes Raman scattering. The effect is exploited by chemists and physicists to gain information about materials for a variety of purposes by performing various forms of Raman spectroscopy. Many other variants of Raman spectroscopy allow rotational energy to be examined and electronic energy levels may be examined if an X-ray source is used in addition to other possibilities. More complex techniques involving pulsed lasers, multiple laser beams and so on are known.

<span class="mw-page-title-main">Two-photon excitation microscopy</span> Fluorescence imaging technique

Two-photon excitation microscopy is a fluorescence imaging technique that is particularly well-suited to image scattering living tissue of up to about one millimeter in thickness. Unlike traditional fluorescence microscopy, where the excitation wavelength is shorter than the emission wavelength, two-photon excitation requires simultaneous excitation by two photons with longer wavelength than the emitted light. The laser is focused onto a specific location in the tissue and scanned across the sample to sequentially produce the image. Due to the non-linearity of two-photon excitation, mainly fluorophores in the micrometer-sized focus of the laser beam are excited, which results in the spatial resolution of the image. This contrasts with confocal microscopy, where the spatial resolution is produced by the interaction of excitation focus and the confined detection with a pinhole.

Coherent anti-Stokes Raman spectroscopy, also called Coherent anti-Stokes Raman scattering spectroscopy (CARS), is a form of spectroscopy used primarily in chemistry, physics and related fields. It is sensitive to the same vibrational signatures of molecules as seen in Raman spectroscopy, typically the nuclear vibrations of chemical bonds. Unlike Raman spectroscopy, CARS employs multiple photons to address the molecular vibrations, and produces a coherent signal. As a result, CARS is orders of magnitude stronger than spontaneous Raman emission. CARS is a third-order nonlinear optical process involving three laser beams: a pump beam of frequency ωp, a Stokes beam of frequency ωS and a probe beam at frequency ωpr. These beams interact with the sample and generate a coherent optical signal at the anti-Stokes frequency (ωprpS). The latter is resonantly enhanced when the frequency difference between the pump and the Stokes beams (ωpS) coincides with the frequency of a Raman resonance, which is the basis of the technique's intrinsic vibrational contrast mechanism.

<span class="mw-page-title-main">Surface-enhanced Raman spectroscopy</span>

Surface-enhanced Raman spectroscopy or surface-enhanced Raman scattering (SERS) is a surface-sensitive technique that enhances Raman scattering by molecules adsorbed on rough metal surfaces or by nanostructures such as plasmonic-magnetic silica nanotubes. The enhancement factor can be as much as 1010 to 1011, which means the technique may detect single molecules.

<span class="mw-page-title-main">Two-photon absorption</span> Simultaneous absorption of two photons by a molecule

In atomic physics, two-photon absorption, also called two-photon excitation or non-linear absorption, is the simultaneous absorption of two photons of identical or different frequencies in order to excite a molecule from one state to a higher energy, most commonly an excited electronic state. Absorption of two photons with different frequencies is called non-degenerate two-photon absorption. Since TPA depends on the simultaneous absorption of two photons, the probability of TPA is proportional to the square of the light intensity; thus it is a nonlinear optical process. The energy difference between the involved lower and upper states of the molecule is equal or smaller than the sum of the photon energies of the two photons absorbed. Two-photon absorption is a third-order process, with absorption cross section typically several orders of magnitude smaller than one-photon absorption cross section.

Chemical imaging is the analytical capability to create a visual image of components distribution from simultaneous measurement of spectra and spatial, time information. Hyperspectral imaging measures contiguous spectral bands, as opposed to multispectral imaging which measures spaced spectral bands.

<span class="mw-page-title-main">Spatially offset Raman spectroscopy</span>

Spatially offset Raman spectroscopy (SORS) is a variant of Raman spectroscopy that allows highly accurate chemical analysis of objects beneath obscuring surfaces, such as tissue, coatings and bottles. Examples of uses include analysis of: bone beneath skin, tablets inside plastic bottles, explosives inside containers and counterfeit tablets inside blister packs. There have also been advancements in the development of deep non-invasive medical diagnosis using SORS with the hopes of being able to detect breast tumors.

<span class="mw-page-title-main">Raman microscope</span> Laser microscope used for Raman spectroscopy

The Raman microscope is a laser-based microscopic device used to perform Raman spectroscopy. The term MOLE is used to refer to the Raman-based microprobe. The technique used is named after C. V. Raman, who discovered the scattering properties in liquids.

<span class="mw-page-title-main">Resonant inelastic X-ray scattering</span> Advanced X-ray spectroscopy technique

Resonant inelastic X-ray scattering (RIXS) is an advanced X-ray spectroscopy technique.

The technique of vibrational analysis with scanning probe microscopy allows probing vibrational properties of materials at the submicrometer scale, and even of individual molecules. This is accomplished by integrating scanning probe microscopy (SPM) and vibrational spectroscopy. This combination allows for much higher spatial resolution than can be achieved with conventional Raman/FTIR instrumentation. The technique is also nondestructive, requires non-extensive sample preparation, and provides more contrast such as intensity contrast, polarization contrast and wavelength contrast, as well as providing specific chemical information and topography images simultaneously.

Stimulated Raman spectroscopy, also referred to as stimulated Raman scattering (SRS) is a form of spectroscopy employed in physics, chemistry, biology, and other fields. The basic mechanism resembles that of spontaneous Raman spectroscopy: a pump photon, of the angular frequency , which is scattered by a molecule has some small probability of inducing some vibrational transition, as opposed to inducing a simple Rayleigh transition. This makes the molecule emit a photon at a shifted frequency. However, SRS, as opposed to spontaneous Raman spectroscopy, is a third-order non-linear phenomenon involving a second photon—the Stokes photon of angular frequency —which stimulates a specific transition. When the difference in frequency between both photons resembles that of a specific vibrational transition the occurrence of this transition is resonantly enhanced. In SRS, the signal is equivalent to changes in the intensity of the pump and Stokes beams. The signals are typically rather low, of the order of a part in 10^5, thus calling for modulation-transfer techniques: one beam is modulated in amplitude and the signal is detected on the other beam via a lock-in amplifier. Employing a pump laser beam of a constant frequency and a Stokes laser beam of a scanned frequency allows for the unraveling of the spectral fingerprint of the molecule. This spectral fingerprint differs from those obtained by other spectroscopy methods such as Rayleigh scattering as the Raman transitions confer to different exclusion rules than those that apply for Rayleigh transitions.

Micro-spatially offset Raman spectroscopy (micro-SORS) is an analytical technique developed in 2014 that combines SORS with microscopy. The technique derives its sublayer‐resolving properties from its parent technique SORS. The main difference between SORS and micro-SORS is the spatial resolution: while SORS is suited to the analysis of millimetric layers, micro-SORS is able to resolve thin, micrometric-scale layers. Similarly to SORS technique, micro-SORS is able to preferentially collect the Raman photons generated under the surface in turbid media. In this way, it is possible to reconstruct the chemical makeup of micrometric multi-layered turbid system in a non destructive way. Micro-SORS is particularly useful when dealing with precious or unique objects as for Cultural Heritage field and Forensic Science or in biomedical applications, where a non-destructive molecular characterization constitute a great advantage.

<span class="mw-page-title-main">Coherent Raman scattering microscopy</span>

Coherent Raman scattering (CRS) microscopy is a multi-photon microscopy technique based on Raman-active vibrational modes of molecules. The two major techniques in CRS microscopy are stimulated Raman scattering (SRS) and coherent anti-Stokes Raman scattering (CARS). SRS and CARS were theoretically predicted and experimentally realized in the 1960s. In 1982 the first CARS microscope was demonstrated. In 1999, CARS microscopy using a collinear geometry and high numerical aperture objective were developed in Xiaoliang Sunney Xie's lab at Harvard University. This advancement made the technique more compatible with modern laser scanning microscopes. Since then, CRS's popularity in biomedical research started to grow. CRS is mainly used to image lipid, protein, and other bio-molecules in live or fixed cells or tissues without labeling or staining. CRS can also be used to image samples labeled with Raman tags, which can avoid interference from other molecules and normally allows for stronger CRS signals than would normally be obtained for common biomolecules. CRS also finds application in other fields, such as material science and environmental science.

Raman spectroelectrochemistry (Raman-SEC) is a technique that studies the inelastic scattering or Raman scattering of monochromatic light related to chemical compounds involved in an electrode process. This technique provides information about vibrational energy transitions of molecules, using a monochromatic light source, usually from a laser that belongs to the UV, Vis or NIR region. Raman spectroelectrochemistry provides specific information about structural changes, composition and orientation of the molecules on the electrode surface involved in an electrochemical reaction, being the Raman spectra registered a real fingerprint of the compounds.

References

  1. Strommen, Dennis P.; Nakamoto, Kazuo (1977). "Resonance raman spectroscopy". Journal of Chemical Education. 54 (8): 474. Bibcode:1977JChEd..54..474S. doi:10.1021/ed054p474. ISSN   0021-9584.
  2. Drago, R.S. (1977). Physical Methods in Chemistry. Saunders. p. 152.
  3. 1 2 3 4 Morris, Michael D.; Wallan, David J. (1979). "Resonance raman spectroscopy: Current applications and prospects". Analytical Chemistry. 51 (2): 182A–192A. doi:10.1021/ac50038a001. ISSN   0003-2700.
  4. Hu, Songzhou; Smith, Kevin M.; Spiro, Thomas G. (January 1996). "Assignment of Protoheme Resonance Raman Spectrum by Heme Labeling in Myoglobin". Journal of the American Chemical Society. 118 (50): 12638–46. doi:10.1021/ja962239e.
  5. 1 2 Clark, Robin J.H.; Dines, Trevor J. (February 1986). "Resonance Raman spectroscopy, and its application to inorganic chemistry". Angewandte Chemie International Edition. 25 (2): 131–158. doi:10.1002/anie.198601311. ISSN   0570-0833.
  6. Austin, J.C.; Rodgers, K.R.; Spiro, T.G. (1993). Protein Structure from ultraviolet resonance Raman spectroscopy. Methods in Enzymology. Vol. 226. pp. 374–396. doi:10.1016/0076-6879(93)26017-4. PMID   8277873.
  7. 1 2 Spiro, T.G.; Czernuszewicz, R.S. (1995). Resonance Raman spectroscopy of metalloproteins. Methods in Enzymology. Vol. 246. pp. 416–460. doi:10.1016/0076-6879(95)46020-9. ISSN   0076-6879. PMID   7752933.
  8. 1 2 3 4 5 6 7 8 9 10 Efremov, Evtem V.; Ariese, Freek; Gooijer, Cees (2008). "Achievements in resonance Raman spectroscopy: Review of a technique with a distinct analytical chemistry potential". Analytica Chimica Acta. 606 (2): 119–134. doi:10.1016/j.aca.2007.11.006. PMID   18082644.
  9. Orlando, Andrea; Franceschini, Filippo; Muscas, Cristian; Pidkova, Solomiya; Bartoli, Mattia; Rovere, Massimo; Tagliaferro, Alberto (2021). "A comprehensive review on Raman spectroscopy applications". Chemosensors. 9 (9): 262. doi: 10.3390/chemosensors9090262 .
  10. Hirakawa, Akiko Y.; Tsuboi, Masamichi (1975). "Molecular geometry in an excited electronic state and a preresonance Raman effect". Science. 188 (4186): 359–361. doi:10.1126/science.188.4186.359. JSTOR   1739341. PMID   17807877. S2CID   7686714.
  11. 1 2 Spiro, T.G.; Stein, Paul (1977). "Resonance effects in vibrational scattering from complex molecules". Annual Review of Physical Chemistry. 28: 501–521. doi:10.1146/annurev.pc.28.100177.002441.
  12. 1 2 Buhrke, David; Hildebrandt, Peter (2020). "Probing structure and reaction dynamics of proteins using time-resolved Raman spectroscopy". Chemical Reviews. 120 (7): 3577–3630. doi:10.1021/acs.chemrev.9b00429. ISSN   0009-2665. PMID   31814387. S2CID   208954659.
  13. Sahoo, Sangram Keshari; Umapathy, Siva; Parker, Anthony W. (2011). "Time-resolved resonance Raman spectroscopy: Exploring reactive intermediates". Applied Spectroscopy. 65 (10): 1087–1115. doi:10.1366/11-06406. ISSN   0003-7028. PMID   21986070. S2CID   20448809.
  14. Spiro, Thomas G. (1985). "Resonance Raman spectroscopy as a probe of heme protein structure and dynamics". Advances in Protein Chemistry. 37: 111–159. doi:10.1016/S0065-3233(08)60064-9. ISBN   9780120342372. ISSN   0065-3233. PMID   2998161.
  15. Mizutani, Yasuhisa (2017). "Time-resolved resonance Raman spectroscopy and application to studies on ultrafast protein dynamics". Bulletin of the Chemical Society of Japan. 90 (12): 1344–1371. doi: 10.1246/bcsj.20170218 . ISSN   0009-2673.
  16. Kelley, Anne Myers (2010). "Hyper-Raman Scattering by Molecular Vibrations". Annual Review of Physical Chemistry. 61 (1): 41–61. doi:10.1146/annurev.physchem.012809.103347. ISSN   0066-426X. PMID   20055673.
  17. Smith, W.E. (2008). "Practical understanding and use of surface-enhanced Raman scattering/surface-enhanced resonance Raman scattering in chemical and biological analysis". Chemical Society Reviews. 37 (5): 955–964. doi:10.1039/b708841h. ISSN   1460-4744. PMID   18443681.
  18. Vogt, Frederick G.; Strohmeier, Mark (2013). "Confocal UV and resonance Raman microscopic imaging of pharmaceutical products". Molecular Pharmaceutics. 10 (11): 4216–4228. doi:10.1021/mp400314s. ISSN   1543-8384. PMID   24050305.
  19. Scarmo, Stephanie; Cartmel, Brenda; Lin, Haiqun; Leffell, David J.; Ermakov, Igor V.; Gellermann, Werner; Bernstein, Paul S.; Mayne, Susan T. (2013). "Single v. multiple measures of skin carotenoids by resonance Raman spectroscopy as a biomarker of usual carotenoid status". British Journal of Nutrition. 110 (5): 911–917. doi:10.1017/S000711451200582X. PMC   3696054 . PMID   23351238.
  20. Senak, L.; Ju, Z.M.; Noy, N.; Callender, R.; Manor, D. (1997). "The interactions between cellular retinol-binding protein (CRBP-I) and retinal: A vibrational spectroscopic study". Biospectroscopy. 3 (2): 131–142. doi:10.1002/(SICI)1520-6343(1997)3:2<131::AID-BSPY6>3.0.CO;2-A. ISSN   1075-4261.
  21. Mathies, Guinevere; van Hemert, Marc C.; Gast, Peter; Gupta, Karthick B. Sai Sankar; Frank, Harry A.; Lugtenburg, Johan; Groenen, Edgar J.J. (2011). "Configuration of spheroidene in the photosynthetic reaction center of Rhodobacter spheroides: A comparison of wild-type and reconstituted R26". Journal of Physical Chemistry A. 115 (34): 9552–9556. doi:10.1021/jp112413d. hdl: 1887/3570972 . ISSN   1089-5639. PMID   21604722.
  22. Stanley, R.J. (2001). "Advances in flavin and flavoprotein optical spectroscopy". Antioxidants and Redox Signaling. 3 (5): 847–866. doi:10.1089/15230860152665028. ISSN   1523-0864. PMID   11761332.
  23. Hirota, S.; Ogura, T.; Appelman, E.H.; Shinzawaitoh, K.; Yoshikawa, S.; Kitagawa, T. (1994). "Observation of a new oxygen-isotope-sensitive Raman band for oxyhemoproteins and its implication in heme pocket structures". Journal of the American Chemical Society. 116 (23): 10564–10570. doi:10.1021/ja00102a025. ISSN   0002-7863.
  24. Mukherjee, Manjistha; Dey, Abhishek (2021). "Rejigging Electron and Proton Transfer to Transition between Dioxygenase, Monooxygenase, Peroxygenase, and Oxygen Reduction Activity: Insights from Bioinspired Constructs of Heme Enzymes". Jacs Au. 1 (9): 1296–1311. doi:10.1021/jacsau.1c00100. ISSN   2691-3704. PMC   8479764 . PMID   34604840.
  25. 1 2 Oladepo, Sulayman A.; Xiong, Kan; Hong, Zhenmin; Asher, Sanford A.; Handen, Joseph; Lednev, Igor K. (2012). "UV resonance Raman investigations of peptide and protein structure and dynamics". Chemical Reviews. 112 (5): 2604–2628. doi:10.1021/cr200198a. PMC   3349015 . PMID   22335827.
  26. Thomas, George J. (1999). "Raman spectroscopy of protein and nucleic acid assemblies". Annual Review of Biophysics and Biomolecular Structure. 28: 1–27. doi:10.1146/annurev.biophys.28.1.1. ISSN   1056-8700. PMID   10410793.
  27. Spencer, Joseph; Nesbitt, John; Trewhitt, Harrison; Kashtiban, Reza; Bell, Gavin; Ivanov, Victor; Faulques, Eric; Smith, David (2014). "Raman Spectroscopy of Optical Transitions and Vibrational Energies of ~1 nm HgTe Extreme Nanowires within Single Walled Carbon Nanotubes" (PDF). ACS Nano. 8 (9): 9044–52. doi:10.1021/nn5023632. PMID   25163005.
  28. Panda, Jaya Kumar; Roy, Anushree; Gemmi, Mauro; Husnau, Elena; Li, Ang; Ercolani, Daniele; Sorba, Lucia (2013). "Electronic band structure of wurtzite GaP nanowires via temperature dependent resonance Raman spectroscopy". Applied Physics Letters. 103 (2): 023108. arXiv: 1303.7058 . doi:10.1063/1.4813625. ISSN   0003-6951. S2CID   93629086.

Further reading