首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
Time-resolved resonance Raman spectroscopy is used to obtain chromophore vibrational spectra of the pR, pB', and pB intermediates during the photocycle of photoactive yellow protein. In the pR spectrum, the C8-C9 stretching mode at 998 cm(-1) is approximately 60 cm(-1) lower than in the dark state, and the combination of C-O stretching and C7H=C8H bending at 1283 cm(-1) is insensitive to D2O substitution. These results indicate that pR has a deprotonated, cis chromophore structure and that the hydrogen bonding to the chromophore phenolate oxygen is preserved and strengthened in the early photoproduct. However, the intense C7H=C8H hydrogen out-of-plane (HOOP) mode at 979 cm(-1) suggests that the chromophore in pR is distorted at the vinyl and adjacent C8-C9 bonds. The formation of pB' involves chromophore protonation based on the protonation state marker at 1174 cm(-1) and on the sensitivity of the COH bending at 1148 cm(-1) as well as the combined C-OH stretching and C7H=C8H bending mode at 1252 cm(-1) to D2O substitution. The hydrogen out-of-plane Raman intensity at 985 cm(-1) significantly decreases in pB', suggesting that the pR-to-pB' transition is the stage where the stored photon energy is transferred from the distorted chromophore to the protein, producing a more relaxed pB' chromophore structure. The C=O stretching mode downshifts from 1660 to 1651 cm(-1) in the pB'-to-pB transition, indicating the reformation of a hydrogen bond to the carbonyl oxygen. Based on reported x-ray data, this suggests that the chromophore ring flips during the transition from pB' to pB. These results confirm the existence and importance of the pB' intermediate in photoactive yellow protein receptor activation.  相似文献   

2.
Iwata T  Nozaki D  Sato Y  Sato K  Nishina Y  Shiga K  Tokutomi S  Kandori H 《Biochemistry》2006,45(51):15384-15391
Phototropin, a blue-light photoreceptor in plants, has two FMN-binding domains named LOV1 and LOV2. We previously observed temperature-dependent FTIR spectral changes in the C=O stretching region (amide-I vibrational region of the peptide backbone) for the LOV2 domain of Adiantum phytochrome3 (phy3-LOV2), suggesting progressive structural changes in the protein moiety (Iwata, T., Nozaki, D., Tokutomi, S., Kagawa, T., Wada, M., and Kandori, H. (2003) Biochemistry 42, 8183-8191). Because FMN also possesses two C=O groups, in this article, we aimed at assigning C=O stretching vibrations of the FMN and protein by using 13C-labeling. We assigned the C(4)=O and C(2)=O stretching vibrations of FMN by using [4,10a-13C2] and [2-13C] FMNs, respectively, whereas C=O stretching vibrations of amide-I were assigned by using 13C-labeling of protein. We found that both C(4)=O and C(2)=O stretching vibrations shift to higher frequencies upon the formation of S390 at 77-295 K, suggesting that the hydrogen bonds of the C=O groups are weakened by adduct formation. Adduct formation presumably relocates the FMN chromophore apart from its hydrogen-bonding donors. Temperature-dependent amide-I bands are unequivocally assigned by separating the chromophore bands. The hydrogen bond of the peptide backbone in the loop region is weakened upon S390 formation at low temperatures, while being strengthened at room temperature. The hydrogen bond of the peptide backbone in the alpha-helix is weakened regardless of temperature. On the other hand, structural perturbation of the beta-sheet is observed only at room temperature, where the hydrogen bond is strengthened. Light-signal transduction by phy3-LOV2 must be achieved by the progressive protein structural changes initiated by the adduct formation of the FMN.  相似文献   

3.
Mid-infrared spectral changes associated with the photoreduction of the bacteriopheophytin electron acceptor H(A) in reaction centers (RCs) of the filamentous anoxygenic phototrophic bacterium Chloroflexus (Cfl.) aurantiacus are examined by light-induced Fourier transform infrared (FTIR) spectroscopy. The light-induced H(A)(-)/H(A) FTIR (1800-1200cm(-1)) difference spectrum of Cfl. aurantiacus RCs is compared to that of the previously well characterized purple bacterium Rhodobacter (Rba.) sphaeroides RCs. The most notable feature is that the large negative IR band at 1674cm(-1) in Rba. sphaeroides R-26, attributable to the loss of the absorption of the 13(1)-keto carbonyl of H(A) upon the radical anion H(A)(-) formation, exhibits only a very minor upshift to 1675cm(-1) in Cfl. aurantiacus. In contrast, the absorption band of the 131-keto C=O of H(A)(-) is strongly upshifted in the spectrum of Cfl. aurantiacus compared to that of Rba. sphaeroides (from 1588 to 1623cm(-1)). The data are discussed in terms of: (i) replacing the glutamic acid at L104 in Rba. sphaeroides R-26 RCs by a weaker hydrogen bond donor, a glutamine, at the equivalent position L143 in Cfl. aurantiacus RCs; (ii) a strengthening of the hydrogen-bonding interaction of the 131-keto C=O of H(A) with Glu L104 and Gln L143 upon H(A)(-) formation and (iii) a possible influence of the protein dielectric environment on the 131-keto C=O stretching frequency of neutral H(A). A conformational heterogeneity of the 133-ester C=O group of H(A) is detected for Cfl. aurantiacus RCs similar to what has been previously described for purple bacterial RCs.  相似文献   

4.
Pastrana-Rios B 《Biochemistry》2001,40(31):9074-9081
Synthetic model helical peptides, Acetyl-W(EAAAR)(5)A-amide with (13)C=O specifically labeled alanine segments in repeats n = 1,2 or 4,5 were studied in aqueous D(2)O solution as a function of temperature using Fourier transform infrared spectroscopy and two-dimensional correlation analysis. The (13)C==O provided a probe which was sensitive to the carbonyl stretch in the peptide bonds of the alanine residues at the amino terminal end in one peptide as compared to the probe in the carboxy terminal end of the other peptide during thermal perturbation. The relative stability of each terminal end was examined; the more stable terminal was determined to be the amino terminal end. Also studied were the glutamate and arginine side-chain modes involved in the salt bridging interaction. Two-dimensional correlation analysis enabled enhanced resolution in the spectral region of 1520--1700 cm(-1), and thus, the order in which these vibrational modes were perturbed as a function of increasing temperature were established.  相似文献   

5.
Shibata M  Muneda N  Sasaki T  Shimono K  Kamo N  Demura M  Kandori H 《Biochemistry》2005,44(37):12279-12286
Halorhodopsin is a light-driven chloride ion pump. Chloride ion is bound in the Schiff base region of the retinal chromophore, and unidirectional chloride transport is probably enforced by the specific hydrogen-bonding interaction with the protonated Schiff base and internal water molecules. In this article, we study hydrogen-bonding alterations of the Schiff base and water molecules in halorhodopsin of Natronobacterium pharaonis (pHR) by assigning their N-D and O-D stretching vibrations in D(2)O, respectively. Highly accurate low-temperature Fourier transform infrared spectroscopy revealed that hydrogen bonds of the Schiff base and water molecules are weak in the unphotolyzed state, whereas they are strengthened upon retinal photoisomerization. Halide dependence of the stretching vibrations enabled us to conclude that the Schiff base forms a direct hydrogen bond with Cl(-) only in the K intermediate. Hydrogen bond of the Schiff base is further strengthened in the L(1) intermediate, whereas the halide dependence revealed that the acceptor is not Cl(-), but presumably a water molecule. Thus, it is concluded that the hydrogen-bonding interaction between the Schiff base and Cl(-) is not a driving force of the motion of Cl(-). Rather, the removal of its hydrogen bonds with the Schiff base and water(s) makes the environment around Cl(-) less polar in the L(1) intermediate, which presumably drives the motion of Cl(-) from its binding site to the cytoplasmic domain.  相似文献   

6.
A vibrational analysis was carried out showing that the infrared experimental data of 13C and 18O carbon monoxide complexes of hemocyanin of Fager and Alben (Biochemistry 11 (1972) 4786) are consistent with a coordination of the carbon atom of CO to one of the two copper ions in the active site. This conclusion contradicts the original interpretation of Fager and Alben in which oxygen-coordination to copper was suggested. This vibrational analysis can also be applied to the study of Alben and Caughey (Biochemistry 7 (1968) 175) with 13C and 18O carbonyl hemoglobin, in which oxygen-coordination to iron was suggested. Carbonyl hemocyanins from several sources have also been studied by infrared spectroscopy. The single stretching vibration of CO bound to arthropodal (Cancer magister) hernocyanin (nu(co)) is at 2042.5 cm(-1), while nu(CO) for gastropod (Helix pomatia of the phylum Mollusca) alpha and beta hemocyanin is at 2064.5 cm(-1)and 2062.5 cm(-1), respectively. The intensities of the CO stretching bands were all around 1.5 X 10(4) M(-1) cm(-2). Calculations show that with the present attainable accuracy it is impossible to detect hydrogen bonding of exchangeable protons to small molecules bound to proteins (for example CO), by comparing its stretching frequencies in H2O and D2O buffers.  相似文献   

7.
The barotropic behavior of neat and aqueous 1,2,3-triacetyl glycerol was investigated by FT-IR spectroscopy over the pressure range 0.001 to 35 kbar. The infrared spectrum in the presence of water shows bands characteristic of hydrogen bonded carbonyl groups. An increase in hydrostatic pressure leads to a strengthening of the intermolecular hydrogen bond between water and the lipid ester C = O groups. The pressure-induced formation of ice VI at 9 kbar does not affect this hydrogen bond, however, the formation, at 20 kbar, of ice VII in which the water/water hydrogen bonds are stronger than the lipid C = O/water hydrogen bonds, frees the lipid carbonyl groups from the hydrogen-bonding to water.  相似文献   

8.
7,7,8,8-Tetracyanoquinodimethane (TCNQ) was incorporated in fully hydrated liposomes of the following pyrene-containing as well as non-labelled phospholipids: 1-palmitoyl-2-[10-(pyren-1-yl)decanoyl]-sn-glycero-3-phosphatid ylc holine (PPDPC), 1-palmitoyl-2-[10-(pyren-1-yl)decanoyl]-sn-glycero-3-phosphatidyl- rac'- glycerol (rac'-PPDPG), 1-palmitoyl-2-[10-(pyren-1-yl)decanoyl]-sn-glycero-3-phosphatidyl- sn-3'- glycerol (3'-PPDPG), 1-[10-(pyren-1-yl)decanoyl]-2-palmitoyl-sn-glycero-3-phosphatidyl- sn-3'- glycerol (3'-PDPPG), 1-[10-pyren-1-yl)decanoyl]-2-palmitoyl-sn-glycero-3-phosphatidyl-s n-1'- glycerol (1'-PDPPG), 1,2-dipalmitoyl-sn-glycero-3-phosphatidylcholine (DPPC) and 1,2-dipalmitoyl-sn-glycero-3-phosphatidyl-rac'-glycerol (rac'-DPPG). Lyophilized charge-transfer (CT) complexes of TCNQ with phospholipids were examined by Fourier transform infrared photoacoustic spectroscopy (FTIR-PAS). Due to the spectral changes observed in the vibrational bands originating from the CH2 and C = O stretching vibrations, and the bands associated with the polar headgroup of the phospholipids it is evident that TCNQ has only a minor perturbing effect on the hydrocarbon chains. However, the molecular interaction between TCNQ and phospholipids is seen in the polar headgroup region. The donated electrons are most likely located on the oxygens of the phosphate group in the polar head. As judged from the present infrared data interactions of TCNQ with phosphatidylcholines (PC) and phosphatidylglycerols (PG) differ. For PG the complex formation produces a second strong C = O stretching band at approx. 1710 cm-1 in addition to the band at approx. 1735 cm-1 indicating a specific molecular interaction in the interfacial region.  相似文献   

9.
Doll C  Bell AF  Power N  Tonge PJ  Tipton PA 《Biochemistry》2005,44(34):11440-11446
The binding of the inhibitor 8-nitroxanthine to urate oxidase has been investigated by Raman and UV-visible absorption spectroscopy. The absorption maximum of 8-nitroxanthine shifts from 380 to 400 nm upon binding to the enzyme, demonstrating that the electronic structure of the ligand is perturbed. It has been proposed that oxidation of the substrate urate by urate oxidase is facilitated by formation of the substrate dianion at the enzyme active site, and Raman spectra of urate oxidase-bound 8-nitroxanthine suggest that both the dianionic and monoanionic forms of the ligand are bound to the enzyme under conditions where in solution the monoanion is present exclusively. The C4-C5 stretching frequency appears as a relatively isolated vibrational mode in 8-nitroxanthine whose frequency shifts according to the protonation state of the purine ring. Identification of the C4-C5 stretching mode was confirmed using [4-(13)C]-8-nitroxanthine and ab initio calculation of the vibrational modes. Two peaks corresponding to the C4-C5 stretching mode were evident in spectra of enzyme-bound 8-nitroxanthine, at 1541 and 1486 cm(-)(1). The higher frequency peak was assigned to monoanionic 8-nitroxanthine, and the low-frequency peak was assigned to dianionic 8-nitroxanthine. The C4-C5 stretching frequency for free monoanionic 8-nitroxanthine was at 1545 cm(-)(1), indicating that the enzyme polarizes that bond when the ligand is bound. The C4-C5 stretching frequency in dianionic 8-nitroxanthine is also shifted by 4 cm(-)(1) to lower frequency upon binding. For 8-nitroxanthine free in solution, the C4-C5 stretching frequency shifts to lower frequency upon deprotonation, and the absorption maximum in the UV-visible spectrum shifts to higher wavelength. The spectral shifts observed upon binding of 8-nitroxanthine to urate oxidase are consistent with increased anionic character of the ligand, which is expected to promote catalysis in the reaction with the natural substrate urate. In the Raman spectra of 8-nitroxanthine bound to the F179A, F179Y, and K9M mutant proteins, the C4-C5 stretching frequency was not perturbed from its position for the unbound ligand. Both V(max) and V/K were decreased in the mutant enzymes, demonstrating a correlation between the interaction that perturbs the C4-C5 stretching frequency and the catalytic activity of the enzyme. It is suggested that hydrogen-bonding interactions that lead to precise positioning and deprotonation of the substrate are perturbed by the mutations.  相似文献   

10.
Wells TA  Takahashi E  Wraight CA 《Biochemistry》2003,42(14):4064-4074
In the primary quinone (Q(A)) binding site of Rb. sphaeroides reaction centers (RCs), isoleucine M265 is in extensive van der Waals contact with the ubiquinone headgroup. Substitution of threonine or serine for this residue (mutants M265IT and M265IS), but not valine (mutant M265IV), lowers the redox midpoint potential of Q(A) by about 100 mV (Takahashi et al. (2001) Biochemistry 40, 1020-1028). The unexpectedly large effect of the polar substitutions is not due to reorientation of the methoxy groups as similar redox potential changes are seen for these mutants with either ubiquinone or anthraquinone as Q(A). Using FTIR spectroscopy to compare Q(A)(-)/Q(A) IR difference spectra for wild type and the M265 mutant RCs, we found changes in the polar mutants (M265IT and M265IS) in the quinone C[double bond]O and C[double bond]C stretching region (1600-1660 cm(-1)) and in the semiquinone anion band (1440-1490 cm(-1)), as well as in protein modes. Modeling the mutations into the X-ray structure of the wild-type RC indicates that the hydroxyl group of the mutant polar residues, Thr and Ser, is hydrogen bonded to the peptide C[double bond]O of Thr(M261). It is suggested that the mutational effect is exerted through the extended backbone region that includes Ala(M260), the hydrogen bonding partner to the C1 carbonyl of the quinone headgroup. The resulting structural perturbations are likely to include lengthening of the hydrogen bond between the quinone C1[double bond]O and the peptide NH of Ala(M260). Possible origins of the IR spectroscopic and redox potential effects are discussed.  相似文献   

11.
Previous vibrational spectroscopic studies of solid acyl-alkyl and diacyl phosphatidylcholines suggested that the sn1- and sn2-carbonyl stretching modes of 1,2-diacylglycerolipids have different absorption maxima. To address the assignment of sn1- and sn2-carbonyl stretching modes of hydrated 1,2-diacylglycerolipids, aqueous dispersions of 1-palmitoyl-2-hexadecyl phosphatidylcholine (PHPC), 1-hexadecyl-2-palmitoyl phosphatidylcholine (HPPC), 1,2-dipalmitoylphosphatidylcholine (DPPC), as well as hydrated samples of unlabeled, sn1-13C=O-labeled, sn2-13C=O-labeled, and doubly 13C=O-labeled dimyristoylphosphatidylcholine (DMPC) were examined by Fourier transform infrared spectroscopy. The ester carbonyl stretching (nu C=O) bands of HPPC and PHPC each exhibit maxima near 1726 cm-1 and appear to be a summation of three subcomponents with maxima near 1740 cm-1, 1725 and 1705-1711 cm-1. In contrast, the nu C=O band of DPPC exhibits its maximum near 1733 cm-1 and appears to be a summation of two components centered near 1742 and 1727 cm-1. Thus the ester carbonyl group of the acyl-alkyl PCs appears to reside in a more polar environment than the ester carbonyl groups of their diacyl analogue. This observation implies that the polar/apolar interfaces of hydrated bilayers formed by PHPC and by HPPC are significantly different from that of DPPC and raises the question of whether the acyl-alkyl PCs are suitable models of their diacyl analogue. The absorption maximum of the nu C=O band of the doubly 13C=O-labeled DMPC occurs near 1691 cm-1 and those of its subcomponents occur near 1699 and 1685 cm-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
Mechanical stretching of covalent bonds, for example when a fibrous polymer is loaded in tension, results in their stretching vibrational bands in the infrared or Raman spectrum being shifted to lower frequency. Conversely stretching a hydrogen bond shifts the stretching vibrational mode of the donor covalent X-H bond to higher frequency. These band shifts are small and difficult to detect in complex regions of the spectrum where differently affected bands overlap. This paper describes a method of integrating the difference spectra (spectrum under tensile strain minus spectrum at zero tensile strain) to recover the shape of the bands that are shifted and the spectral variation in bandshift. The application of this method to two sets of vibrational spectra of cellulose under tension is described. In one example, C-O-C stretching bands of highly crystalline tunicate cellulose were observed to shift to lower frequency under axial strain. In the other example, a group of overlapping O-D stretching bands in partially deuterated cellulose showed varied bandshifts under axial strain, some bandshifts being positive as expected due to extension of axially oriented hydrogen bonds while others were negative. The possibility of constructing spectral plots of bandshift has the potential to clarify the interpretation of overlapped, shifting bands in the vibrational spectra of polymers under tension.  相似文献   

13.
Developing applications for metal-mediated base pairs (metallo-base-pair) has recently become a high-priority area in nucleic acid research, and physicochemical analyses are important for designing and fine-tuning molecular devices using metallo-base-pairs. In this study, we characterized the Hg(II)-mediated T-T (T-Hg(II)-T) base pair by Raman spectroscopy, which revealed the unique physical and chemical properties of Hg(II). A characteristic Raman marker band at 1586 cm(-1) was observed and assigned to the C4=O4 stretching mode. We confirmed the assignment by the isotopic shift ((18)O-labeling at O4) and density functional theory (DFT) calculations. The unusually low wavenumber of the C4=O4 stretching suggested that the bond order of the C4=O4 bond reduced from its canonical value. This reduction of the bond order can be explained if the enolate-like structure (N3=C4-O4(-)) is involved as a resonance contributor in the thymine ring of the T-Hg(II)-T pair. This resonance includes the N-Hg(II)-bonded state (Hg(II)-N3-C4=O4) and the N-Hg(II)-dissociated state (Hg(II+) N3=C4-O4(-)), and the latter contributor reduced the bond order of N-Hg(II). Consequently, the Hg(II) nucleus in the T-Hg(II)-T pair exhibited a cationic character. Natural bond orbital (NBO) analysis supports the interpretations of the Raman experiments.  相似文献   

14.
Polarized, low-temperature Fourier transform infrared (FTIR) difference spectroscopy has been used to investigate the structure of bacteriorhodopsin (bR) as it undergoes phototransitions from the light-adapted state, bR570, to the K630 and M412 intermediates. The orientations of specific retinal chromophore and protein groups relative to the membrane plane were calculated from the linear dichroism of the infrared bands, which correspond to the vibrational modes of those groups. The linear dichroism of the chromophore C=C and C-C stretching modes indicates that the long axis of the polyene chain is oriented at 20-25 degrees from the membrane plane at 250 K and that it orients more in-plane when the temperature is reduced to 81 K. The polyene plane is found to be approximately perpendicular to the membrane plane from the linear dichroism calculations of the HOOP (hydrogen out-of-plane) wags. The orientation of the transition dipole moments of chromophore vibrations in the K630 and M412 intermediates has been probed, and the dipole moment direction of the C=O bond of an aspartic acid that is protonated in the bR570----M412 transition has been measured.  相似文献   

15.
The physical interactions that occur between the nicotinic acetylcholine receptor from Torpedo and the agonists carbamylcholine and tetramethylamine have been studied using both conventional infrared difference spectroscopy and a novel double-ligand difference technique. The latter was developed to isolate vibrational bands from residues in a membrane receptor that interact with individual functional groups on a small molecule ligand. The binding of either agonist leads to an increase in vibrational intensity at frequencies centered near 1663, 1655, 1547, 1430, and 1059 cm(-1) indicating that both induce a conformational change from the resting to the desensitized state. Vibrational shifts near 1580, 1516, 1455, 1334, and between 1300 and 1400 cm(-1) are assigned to structural perturbations of tyrosine and possibly both tryptophan and charged carboxylic acid residues upon the formation of receptor-quaternary amine interactions, with the relatively intense feature near 1516 cm(-1) indicating a key role for tyrosine. Other vibrational bands suggest the involvement of additional side chains in agonist binding. Two side-chain vibrational shifts from 1668 and 1605 cm(-1) to 1690 and 1620 cm(-1), respectively, could reflect the formation of a hydrogen bond between the ester carbonyl of carbamylcholine and an arginine residue. The results demonstrate the potential of the double-ligand difference technique for dissecting the chemistry of membrane receptor-ligand interactions and provide new insight into the nature of nicotinic receptor-agonist interactions.  相似文献   

16.
Spectroscopic evidence is presented for the lack of intramolecular hydrogen bonding in a simple peptide derivative of 5,5-dimethylthiazolidine-4-carboxylic acid (Dtc). The infrared spectrum of Boc-Pro-Ile-OMe 1 in nonpolar solvents displays two N-H stretching bands at 3419 and 3330 cm-1 in CCl4 and one at 3417 and 3328 cm-1 in CHCl3. The low frequency band at 3328-3330 cm-1 may be assigned to conformations with an intramolecular hydrogen bond between the Ile N-H and Boc C = O. The band at 3417-3419 cm-1 is the normal Ile N-H stretch. In the polar solvent CH3CN only one NH stretching band at 3365 cm-1 is observed. The IR spectrum of Boc-Dtc-Ile-OMe 2, on the other hand, displays one N-H stretching band at 3423 cm-1 in CCl4 and one at 3418 cm-1 in CHCl3. The IR spectrum of 2 does not display the N-H stretching band that would arise from intramolecular hydrogen bonding between the Boc C = O and Ile N-H. The lack of intramolecular hydrogen bonding for Boc-Dtc-Ile-OMe 2 was evident also in the NMR spectra in nonpolar solvents. The 1H-NMR spectrum of the Pro dipeptide 1 in 50% CDCl3/C6D6 at 20 degrees displayed two Ile-NH signals at 6.58 and 7.74 ppm. The latter signal corresponds to the intramolecularly hydrogen bonded Ile-NH in the trans-Boc isomer of 1 (60% of the total population), while the former signal corresponds to the nonhydrogen bonded Ile-NH in the cis-Boc isomer.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

17.
Peng Y  Wu P  Siesler HW 《Biomacromolecules》2003,4(4):1041-1044
In the present contribution, two-dimensional ATR-FTIR spectroscopy was used to study the diffusion of water in poly(epsilon-caprolactone) (PCL). In the spectral region of the nu(OH) stretching vibration of water, four absorption bands (3635, 3560, 3411, and 3220 cm(-1)) can be identified. The higher wavenumber band pair at 3635 and 3560 cm(-1) is assigned to the antisymmetric and symmetric OH stretching vibrations, respectively, of water which is partially hydrogen-bonded to the carbonyl groups of PCL, whereas the lower frequency band pair at 3411 (antisymmetric) and 3220 cm(-1) (symmetric) is attributed to the OH stretching vibrations of bulk water which is fully hydrogen-bonded to other water molecules. From the asynchronous map of a 2D correlation analysis of spectra recorded during the diffusion of water into PCL, it was concluded that during the diffusion process the water molecules first penetrate into the free volume (microvoids) of the PCL matrix or are molecularly dispersed in the polymer matrix and then form hydrogen bonds with the C=O groups of the polymer.  相似文献   

18.
T Xiang  D J Goss    M Diem 《Biophysical journal》1993,65(3):1255-1261
We report observed and computed infrared (vibrational) circular dichroism spectra of a number of polyribonucleic acids in aqueous solutions in the 1600-1750 cm-1 spectral region, in which C = O and some nucleotide base ring stretching vibrations occur. The experimental data are compared with results calculated using different levels of sophistication within the exciton approach. We find that observed band shapes are generally well reproduced by these models, particularly if care is taken to determine the direction of the vibrational dipole transition moments accurately.  相似文献   

19.
The identification of structural markers for B12/protein interactions is crucial to a complete understanding of vitamin B12 transport and metabolic reaction mechanisms of B12 coenzymes. Fourier transform infrared spectroscopy can provide direct measurements of changes in the side chains and corrin ring resulting from B12/protein interactions. Using FTIR spectroscopy in various solvent systems, we have identified structural markers for corrinoids in the physiological state. We assign the major band (denoted B), which occurs at ca. 1630 cm-1 in D2O and ca. 1675 cm-1 in ethanol, to the amide I C=O stretching mode of the propionamide side chains of the corrin ring. The lower frequency of band B in D2O versus ethanol is due to the greater hydrogen-bonding properties of D2O that stabilize the charged amide resonance form. Since the propionamides are known to be important in protein binding, band B is a suitable marker for monitoring the interaction of these side chains with proteins. We assign bands at ca. 1575 and 1545 cm-1 (denoted C and D) as breathing modes of the corrin ring on the basis of the bands' solvent independence and their sensitivity to changes in axial ligation. As the sigma-donating strength of the axial ligands increases, the frequencies of bands C and D decrease, possibly indicating a lengthening of the corrin conjugated system. Band A, the known cyanide stretching frequency at ca. 2130 cm-1, probes the cobalt-carbon distance in cyanocorrinoids. As the frequency of band A increases, the cobalt-carbon bond strength should decrease.  相似文献   

20.
A Blume  W Hübner  G Messner 《Biochemistry》1988,27(21):8239-8249
Fourier transform infrared spectroscopy has been used to characterize the carbonyl stretching vibration of DMPC, DMPE, DMPG, and DMPA, all labeled with 13C at the carbonyl group of the sn-2 chain. Due to the vibrational isotope effect, the 13C = O and the 12C = O vibrational bands are separated by ca. 40-43 cm-1. This frequency difference does not change when the labeling is reversed with the 13C = O group at the sn-1 chain. For lipids in organic solvents possible conformational differences between the sn-1 and sn-2 ester groups have no effect on the vibrational frequency of the C = O groups. In aqueous dispersion unlabeled phospholipids always show a superposition of two bands for the C = O vibration located at ca. 1740 and 1727 cm-1. These two bands have previously been assigned to the sn-1 and sn-2 C = O groups. FT-IR spectra of 13C-labeled phospholipids show that the vibrational bands of both, the sn-1 as well as the sn-2 C = O group, are clearly superpositions of at least two underlying components of different frequency and intensity. Band frequencies were determined by Fourier self-deconvolution and second-derivative spectroscopy. The difference between the component bands is ca. 11-17 cm-1. Again, the conformational effect as shown by reversed labeling is negligible with only 1-2 cm-1. The splitting of the C = O vibrational bands in H2O and D2O is caused by hydrogen bonding of water molecules to both C = O groups as shown by a comparison with spectra of model ester compounds in different solvents. To extract quantitative information about changes in hydration, band profiles were stimulated with Gaussian-Lorentzian functions. The chemical nature of the head group and its electronic charge have distinctive effects on the extent of hydration of the carbonyl groups. In the gel and liquid-crystalline phase of DMPC the sn-2 C = O group is more hydrated than the sn-1 C = O. This is accord with the conformation determined by X-ray analysis. In DMPG the sn-1 C = O group seems to be more accessible to water, indicating a different conformation of the glycerol backbone.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号