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 31 毫秒
1.
N-Acetyl-amino acid methylamides CH3CONHCHRCONHCH3 were prepared from L - and DL -alanine, L and DL -α-amino-n-butyric acid, L - and DL -norvaline, DL -norleucine, L - and DL -methionine, L - and DL -leucine, L -aspartic acid and DL -phenylalanine. The deuterium homologs of the type CH3CONDCHRCONDCH3, CD3CONHCHRCONHCH3, and CH3CONHCHRCONHCD3 were also prepared. The infrared spectra of these compounds were measured down to 300 cm?1 in the crystalline state. The infrared spectra of N-isopropylacetamide CH3CONHCH(CH3)2, N-methylisobutyramide (CH3)2CHCONHCH3 and their deuterium homologs were also measured. The C?O in-plane and out-of-plane bending vibration bands of the CH3CONHCα group (amide IVa and VIa) and those of the –CαCONHCH3 group (amide IVb and VIb) were assigned from these data. Two crystalline modifications, form I and form II, were found for the compounds prepared from L -alanine, DL -leucine, L -aspartic acid and DL -phenylalanine. The two forms show quite different skeletal vibrations, which suggest, rotational isomerism. Two distinct patterns were found as to the positions of the amide IVa and VIa bands for the above compounds. The two amide bands were found near 630 and 600 cm?1 in form I, whereas they were found near 600 cm?1 in form II. The crystals of the remaining compounds were also classified into form I or form II on the basis of the arrangement of the amide bands. The X-ray structure analyses suggest that these two forms have different hydrogen-bond structures.  相似文献   

2.
Raman spectra are presented for sarcoplasmic reticulum membranes. Interpretation of the 1000–1130 cm?1 region of the spectrum indicates that the sarcoplasmic reticulum membrane may be more fluid than erythrocyte membranes that have been examined by the same technique. The fluidity of the membrane also manifests itself in the amide I portion of the membrane spectrum with a strong 1658 cm?1 band characteristic of CC stretching in hydrocarbon side chains exhibiting cis conformation. This band is unaltered in intensity and position in H2O and in 2H2O thus obscuring amide I protein conformation. Of particular interest is the appearance of strong, resonantly enhanced bands at 1160 and 1527 cm?1 attributable to membrane-associated carotenoids.  相似文献   

3.
Studies suggested that the pathogenesis of inflammatory breast cancer (IBC) is related to inflammatory manifestations accompanied by specific cellular and molecular mechanisms in the IBC tumor microenvironment (TME). IBC is characterized by significantly higher infiltration of tumor-associated macrophages (TAMs) that contribute to its metastatic process via secreting many cytokines such as TNF, IL-6, IL-8, and IL-10 that enhance invasion and angiogenesis. Thus, there is a need to first understand how IBC-TME modulates the polarization of TAMs to better understand the role of TAMs in IBC. Herein, we used gene expression signature and Synchrotron Fourier-Transform Infrared Microspectroscopy (SR-μFTIR) to study the molecular and biochemical changes, respectively of in vitro polarized TAMs stimulated by the secretome of IBC and non-IBC cells. The gene expression signature showed significant differences in the macrophage's polarization-related genes between stimulated TAMs. FTIR spectra showed absorption bands in the region of 1700–1500 cm?1 attributed to the amide I ν(C=O), & νAS (CN), δ (NH), and amide II ν(CN), δ (NH) proteins bands. Moreover, three peaks of different intensities and areas were detected in the lipid region of the νCH2 and νCH3 stretching modes positioned within the 3000–2800 cm?1 range. The PCA analysis for the second derivative spectra of the amide regions discriminates between stimulated IBC and non-IBC TAMs. This study showed that IBC and non-IBC TMEs differentially modulate the polarization of TAMs and SR-μFTIR can determine these biochemical changes which will help to better understand the potential role of TAMs in IBC.  相似文献   

4.
To investigate molecular effects of 1‐Ethyl‐3‐(3‐dimethylaminopropyl) carbodiimide (EDC), EDC/N‐hydroxysuccinimide (NHS), glyceraldehyde cross‐linking as well as polymerization temperature and concentration on the three‐dimensional (3D) collagen hydrogels, we analyzed the structures in situ by Raman microspectroscopy. The increased intensity of the 814 and 936 cm?1 Raman bands corresponding to the C—C stretch of a protein backbone and a shift in the amide III bands from 1241 cm?1/1268 cm?1 in controls to 1247 cm?1/1283 cm?1 in glyceraldehyde‐treated gels indicated changes to the alignment of the collagen molecules, fibrils/fibers and/or changes to the secondary structure on glyceraldehyde treatment. The increased intensity of 1450 cm?1 band and the appearance of a strong peak at 1468 cm?1 reflected a change in the motion of lysine/arginine CH2 groups. For the EDC‐treated collagen hydrogels, the increased intensity of 823 cm?1 peak corresponding to the C—C stretch of the protein backbone indicated that EDC also changed the packing of collagen molecules. The 23% decrease in the ratio of 1238 cm?1 to 1271 cm?1 amide III band intensities in the EDC‐modified samples compared with the controls indicated changes to the alignment of the collagen molecules/fibrils and/or the secondary structure. A change in the motion of lysine/arginine CH2 groups was detected as well. The addition of NHS did not induce additional Raman shifts compared to the effect of EDC alone with the exception of a 1416 cm?1 band corresponding to a COO? stretch. Overall, the Raman spectra suggest that glyceraldehyde affects the collagen states within 3D hydrogels to a greater extent compared to EDC and the effects of temperature and concentration are minimal and/or not detectable. © 2012 Wiley Periodicals, Inc. Biopolymers 99: 349–356, 2013.  相似文献   

5.
Fourier transform infrared spectra were obtained for mammalian calmodulin and two of its fragments produced by limited proteolysis with trypsin TR1C (1–77) and TR2C (78–148). Experiments were done in H2O, D2O and D2O/trifluoroethanol (TFE) mixtures. Information about secondary structure was obtained from analysis of the amide I and II bands; while characteristic absorbances for tyrosine, phenylalanine and carboxylate groups were analyzed for changes in tertiary structure. Our data indicate that the secondary and tertiary structure is preserved in the two half molecules of CaM, both in the apo- and Ca2+-saturated state. Addition of the structure-inducing solvent TFE causes marked changes only in the apo-TR1C domain. The maximum wavenumber for the amide I band of the two domains of CaM in D20 was markedly different (1642 cm–1 for TR1C versus 1646/1648 cm–1 for Ca 2+ and apo-TR2C). This renders the amide I band for the intact protein very broad in comparison to that in other proteins and is indicative of a distribution of -helices with slightly different hydrogen bonding patterns.  相似文献   

6.
We have investigated the phase characteristics of 1,2-bis(tricosa-10,12-diynoyl)-sn-glycero-3-phosphocholine (DC23PC), a phosphatidylcholine with diacetylenic groups in the acyl chains, and its saturated analog 1,2-ditricosanoyl-sn-glycero-3-phosphocholine (DTPC), using Fourier-transform infrared spectroscopy (FTIR). Previous studies on the phase behavior of DC23PC in H2O have shown that DC23PC exhibits: (1) formation of cylindrical structures (‘tubules’) by cooling fluid phase multilamellar vesicles (MLVs) through Tm (43° C), and 2) metastability of small unilamellar vesicles (SUVs) in the liquid-crystalline state some 40° C below Tm, with subsequent formation of a gel phase comprised of multilamellar sheets at 2° C. The sheets form tubules when heated and cooled through Tm. FTIR results presented here indicate that as metastable SUVs are cooled toward the transition to bilayer sheets, spectroscopic changes occur before the calorimetric transition as measured by a reduction in the CH2 symmetric stretch frequency and bandwidth. In spite of the vastly different morphologies, the sheet gel phase formed from SUVs is spectroscopically similar to the tubule gel phase. The C-H stretch region of DC23PC gel phase shows bands at 2937 and 2810 cm−1 not observed in the saturated analog of DC23PC, which may be related to perturbations in the acyl chains introduced by the diacetylenic moiety. The narrow CH2 scissoring mode at 1470 cm−1 and the prominent CH2 wagging progression indicate that DC23PC gel phase was highly ordered acyl chains with extended regions of all-trans methylene segments. In addition, the 13 cm−1 reduction in the C  O stretch frequency (1733–1720 cm−1) during the induction of DC23PC gel phase indicates that the interfacial region is dehydrated and rigid in the gel phase.  相似文献   

7.
Comparative CD and Fourier transform ir (FTIR) spectroscopic data on N-Boc protected linear peptides with or without the (Pro-Gly) β-turn motif (e.g., Boc-Tyr-Pro-Gly-Phe-Leu-OH and Boc-Tyr-Gly-Pro-Phe-Leu-OH) are reported herein. The CD spectra, reflecting both backbone and aromatic contributions, were not found to be characteristic of the presence of β-turns. In the amide I region of the FTIR spectra, analyzed by self-deconvolution and curve-fitting methods, the β-turn band shewed up between 1639 and 1633 cm?1 in trifluoroethanol (TFE) but only for models containing the (Pro-Gly) core. This band war-also present in the spectra in chloroform but absent in dimethylsulfoxide. These findings, in agreement with recent ir data on cyclic models and 310-helical polypeptides and protein in D2O [see S. J. Prestrelski, D. M. Byler, and M. P. Thompson (1991), International Journal of Peptide and Protein Research, Vol. 37, pp. 508–512; H. H. Mantsch, A. Perczel. M. Hollósi, and G. D. Fasman (1992), FASEB Journal, Vol. 6, p. A341; H. H. Mantsch. A. Perczel, M. Hollósi, and G. Fasman (1992), Biopolymers. Vol. 33, pp. 201–207; S. M. Miick, G. V. Martinez, W. R. Fiori, A. P. Tedd, and G. L. Millhauser (1992). Nature, Vol. 359, pp. 653–655], suggest that the amide I band, with a major contribution from the acceptor C ? O of the 1 ← 4 intramolecular H bond of β-turns, appears near or below 1640 cm?1, rather than above 1660 cm?1. In TFE, bands between 1670 and 1660 cm?1 are mainly due to “free” carbonyls, that is, C ? O's of amides that are solvated but not involved in the characteristic H bonds of periodic secondary structures or β-turns. © 1994 John Wiley & Sons, Inc.  相似文献   

8.
《FEBS letters》1985,187(2):227-232
Fourier transform infrared (FTIR) difference spectroscopy of the primary electron donor (P) photo-oxidation has been performed for reaction centers (RCs) and chromatophores of purple photosynthetic bacteria. In the 1800–650 cm−1 spectral region highly reproducible absorbance changes were obtained that can be related to specific changes of individual bond absorption. Several bands in the difference spectra are tentatively assigned to changes of intensity and position of the keto and ester CO vibrations of the P bacteriochlorophylls, and a possible interpretation in terms of changes of their environment or type of bonding to the protein is given. Small difference bands in the amide I and II region allow only minor protein conformational changes.  相似文献   

9.
Diatoms are photosynthetic unicellular microalgae and are nature’s hidden source of several biosynthetic metabolites with their use in biofuel, food and drug industries. They mainly contain various lipids, sterols, isoprenoids and toxins with their use in apoptotic, fertility controlling and cancer drugs. Chemical studies on diatoms are limited due to various limitations such as variation of nutrients, contaminants and change in seasonal factors in the environment. To overcome these limitations, we obtained axenic cultures of 12 fresh-water diatom strains on the 22nd day of inoculation having a dry weight of 1 mg each and performed their Fourier transform infrared (FTIR) study for the detection of functional groups responsible for their chemical moiety. The spectral mapping showed a varied level of polyunsaturated fatty acids, amides, amines, ketone bodies and esters for their applications in various pharmacological, food and biofuel industries in the exponential phase of their growth in f/2 media. The FTIR study of the 12 diatom strains showed various similarities in the form of some common peak patterns ranging from 3000 to 3600 cm?1 for νO–H absorption. The symmetric stretching vibration frequency of Diadesmis confervaceae (V2) type species showed different behaviour than others in the spectral region starting from 1600 to 1700 cm?1. The absorption between 1500 and 1575 cm?1 reflects the presence of the –N–H group. Infrared (IR) absorptions falling between 1600 and 1700 cm?1 reflect the presence of amide’s νC=O in all species. Placoneis elginensis (V8) type species showed an additional absorption band which is centred around 1735–1750 cm?1 which perhaps reflects the presence of ester’s νC=O. Diadesmis confervaceae (V2), Nitzschia palea (V4), Placoneis elginensis (V8), Nitzschia palea var. debilis (V6), Nitzschia inconspicua (V10), Gomphonema parvulum (V11) and Sellaphora (V12) showed distinct structural features with important key functionalities that can make them essential drug markers in the pharmaceutical industry.  相似文献   

10.
The vibrational Raman optical activity (ROA) spectra of di- and tri-L -alanine in the range 650–1750 cm?1 have been measured in H2O and D2O solution at high, neutral, and low pH and pD. Corresponding ROA spectra for tetra- and penta-L -alanine have also been obtained, but over a more restricted set of pH and pD conditions. There are similarities with the ROA spectrum of L -alanine below ~ 1200 cm?1, but the spectra are very different above this wavenumber due to the influence of the vibrational coordinates of the peptide group. The similar overall appearance of the di-, tri-, and tetrapeptide ROA under selected conditions of pH and pD, and of all four peptide ROA spectra in DCl and HCl solutions, in the backbone skeletal stretch region ~ 1050–1200 cm?1 and the extended amide III region ~ 1250–1350 cm?1, suggests that the backbone conformation is approximately the same in all four structures. One difference, however, is a shift of a large positive ROA band in H2O at ~ 1341 cm?1 in the dipeptide, assigned to Cα–H and in-plane N–H deformations, down to ~ 1331 cm ?1 in the tripeptide and to ~ 1315 cm?1 in the tetrapeptide and pentapeptide (the last in HCl due to insufficient solubility in H2O), which indicates increasing delocalization of the corresponding normal mode with increasing chain length. Our results do not support the suggestion that stabilizing interactions of the zwitterionic end groups in tri-L -alanine at neutral pH leads to a different solution structure to that at high pH. © 1994 John Wiley & Sons, Inc.  相似文献   

11.
A series of meso-5,10,15-tris(N-methyl-4-pyridiniumyl)-20-(4-alkylamidophenyl) porphyrins were synthesized by derivatizing the amino group on the phenyl ring with the following hydrophobic groups: –C(O)C7F15, –C(O)CHCH2, C(O)CH3, –C(O)C7H15, and –C(O)C15H31. The cationic tris-pyridiumyl porphyrin core serves as a DNA binding motif and a photosensitizer to photomodify DNA molecules. The changes of the UV–Vis absorption spectra during the titration of these porphyrins with calf thymus DNA revealed a large bathochromic shift (up to 14 nm) and a hypochromicity (up to 55%) of the porphyrins Soret bands, usually considered as proof of porphyrin intercalation into DNA. Association constants (K) calculated according to the McGhee and von Hippel model, were in the range of 106–107 M−1. An increase in hydrophobicity of the substituents at the 20−meso-position produced higher binding affinity. These porphyrins caused photomodification of the supercoiled plasmid DNA when a green laser beam at 532 nm was applied. Those with higher surface activity acted more efficiently as DNA photomodifiers. The porphyrin with a perfluorinated alkyl chain (–COC7F15) at the meso-20-position inhibited the growth of gram-positive bacteria (S. aureus, or S. epidermidis). Other porphyrins exhibited moderate activity against both gram-negative and gram-positive organisms.  相似文献   

12.
The selectivity of phosphoryl P(O)R3, sulfoxide S(O)R2, and carbonyl C(O)R2 (R?=?NH2, CH3, OH, and F) derivatives with lanthanide cations (La3+, Eu3+, Lu3+) was studied by density functional theory calculations. Theoretical approaches were also used to investigate energy and the nature of metal–ligand interaction in the model complexes. Atoms in molecules and natural bond orbital (NBO) analyses were accomplished to understand the electronic structure of ligands, L, and the related complexes, L–Ln3+. NBO analysis demonstrated that the negative charge on phosphoryl, carbonyl, and sulfoxide oxygen (OP, OC, and OS) has maximum and minimum values when the connected –R groups are –NH2 and –F. The metal–ligand distance declines as, –F?>?–OH?>?–CH3?>?–NH2. Charge density at the bond critical point and on the lanthanide cation in the L–Ln3+ complexes varies in the order –F?<?–OH?<?–CH3?<?–NH2, due to greater ligand to metal charge transfer, which is well explained by energy decomposition analysis. It was also illustrated that E(2) values of Lp(N)?→?σ*(Y–N) vary in the order P=O ? S=O ? C=O and the related values of Lp(N)?→?σ*(Y=O) change as C=O ? S=O ? P=O in (NH2)nYO ligands (Y?=?P, C, and S). Trends in the L–Ln3+ CP–corrected bond energies are in good accordance with the optimized OY?Ln distances. It seems that, comparing the three types of ligands studied, NH2–substituted are the better coordination ligands.
Graphical Abstract Density functional theory (B3LYP) calculations were used to compare structural, electronic and energy aspects of lanthanide (La, Eu, Lu) complexes of phosphine derivatives with those of carbonyls and sulfoxides in which the R– groups connected to the P=O, C=O and S=O are –NH2, –CH3, –OH and –F.
  相似文献   

13.
Fourier transform infrared (FT‐IR) spectroscopy combined with 2D correlation spectroscopy has been used to offer some information about stability and structure of some soluble elastins. Temperature has been chosen as the perturbation to monitor the infrared behavior of various soluble elastins, namely, α‐elastin p, α‐elastin, and k‐elastin. In the 3800–2700 cm?1 region, the H‐containing groups were analyzed. The bonded hydroxyls are found to decrease prior to the NH‐related hydrogen bonds and also to the conformational reorganization of hydrocarbon chains. The transition temperatures were evaluated and they were found to agree with those obtained from DSC data. The FTIR spectra and their 2nd derivatives denote that α‐ elastins exhibited amide‐I, ‐II and ‐III bands at 1656, 1539 and 1236 cm?1, respectively, while in k‐elastin these bands were found at 1652 cm?1 for amide I, 1540 cm?1 for amide II and 1248 cm?1 for amide III. The macroscopic IR finger‐print method, which combines: general IR spectra, secondary derivative spectra, and 2D‐IR correlation spectra, is useful to discriminate different elastins. Thus using the differences of the position and intensity of the bands from “fingerprint region” of studied elastins, which include the peaks assigned to C?O, C? C groups from α‐helix, β‐turn, and the peaks assigned to the amide groups, it is possible to identify and discriminate elastins from each others. Furthermore, the pattern of 2D‐IR correlation spectra under thermal perturbation, allow their direct identification and discrimination. © 2010 Wiley Periodicals, Inc. Biopolymers 93: 1072–1084, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

14.
The silk protein from the web of Orthaga exvinacea was isolated, purified, and casted into films. This film was treated separately with methanol, acetone, ethyl acetate, and isopropyl alcohol in 50 % concentration for about 30 min. The treated films were thus dried in a desiccator and subjected to FTIR and TG-DTA analysis. The structural studies revealed that the organic solvents induce conformatory changes in the protein film, especially the most sensitive amide I (1650 cm?1) band. This band had shifted to lower wavenumber (1633–1636 cm?1). Furthermore, the conformatory characteristics associated with amide I band also changed from random coil to β-sheet. Generally, β-sheet contributes strength to the protein film. Among the treated films, film treated with acetone showed much thermal stability. Moreover, the film treated with methanol had shown two different temperatures of maximum degradation. It is concluded that in addition to β-sheet content, various other factors such as various processing conditions and structural organization of protein may influence the stability of the films.  相似文献   

15.
The light-induced Q A /QA FTIR difference spectra of Rb. sphaeroides and Rp. viridis show very broad positive bands of small amplitude peaking around 2750 cm–1. Upon 1H/2H exchange these bands shift to about 2150 cm–1. Similarly, the Q B /QB spectra exhibit broad continuum bands at 2600 and 2800 cm–1 shifting to 2100 and 2200 cm–1 in 2H2O for Rb. sphaeroides and Rp. viridis, respectively. These continuum bands are tentatively interpreted in terms of highly polarizable hydrogen bonds in a large web of polar bonds involving cofactors, amino acid residues, and structured water molecules. As a working hypothesis, we propose that the protons participating in this web redistribute upon quinone reduction, increasing their concentration around the newly formed charged species, and leading to net proton uptake. Assuming that the precise localization of the mobile protons is dependent on the local electrostatic, this model can explain the apparent discrepancies between some results of FTIR experiments and of electrostatic calculations. Notably, it could help rationalize the observation that mobile protons tend to localize on Glu L212 upon QB reduction in Rb. sphaeroides, while for QB reduction in Rp. viridis and for QA reduction in both Rb. sphaeroides and Rp. viridis, proton uptake by a small number of carboxylic residues is not supported by the FTIR data.  相似文献   

16.
Infrared spectra of poly-L -alanine in trifluoroacetic acid-chloroform mixtures have been investigated and compared with those of a model amide (N-methylacetamide). The purpose of this work is to determine the nature of peptide-acid specific interactions responsible for the helix-random coil transition of polymer chains. Analysis is made in using amide (A, I, II, III) and acid (νC?O, νOH) vibrations which are specially sensitive to molecular interactions. We examined a model compound to determine the spectral characteristics of the different complexes or species formed between amide and acid. At a low acid concentration, hydrogen-bonded complexes: ? (NH) C?O…?HOOCCF3 (1) are evidenced but no association between amide NH and acid CO groups (complexes A) is observed. For higher acid concentrations complexes (I) are progressively changed into ions pairs and free ions, which result from amide protonation by acid, according to the exothermic equilibrium (I)?? (NH)COH+, ?OOCCF3(II). Amidium and carboxylate bands are localized between 1680–1705 cm?1 and 1620–1625 cm?1, respectively. If the cation band is always clearly seen, the anion band is only observed for the most acidic solutions. For the polymer, a gradual complexation of type (I) is observed for all acid concentrations. From our results, the assumption of an (A) type interaction seems very unlikely but cannot be excluded. Moreover, proton transfer—similar to that observed with a model amide—is never evidenced since, in particular, the amidium band characteristic of protonation is never seen. In contrast to previous investigations, we conclude that the helix-random coil transition of polypeptides is not due to the protonation of the peptide functions. This transition does suggest a strong interaction by hydrogen bonds between polymer and acid molecules.  相似文献   

17.
18.
Abstract

The present study aims at evaluating a batch scale biosorption potential of Moringa oleifera leaves (MOL) for the removal of Pb(II) from aqueous solutions. The MOL biomass was characterized by FTIR, SEM, EDX, and BET. The impact of initial concentrations of Pb (II), adsorbent dosage, pH, contact time, coexisting inorganic ions (Ca2+, Na+, K+, Mg2+, CO32?, HCO3?, Cl?), electrical conductivity (EC) and total dissolved salts (TDS) in water was investigated. The results revealed that maximum biosorption (45.83?mg/g) was achieved with adsorbent dosage 0.15?g/100?mL while highest removal (98.6%) was obtained at adsorbent biomass 1.0?g/100?mL and pH 6. The presence of coexisting inorganic ions in water showed a decline in Pb(II) removal (8.5% and 5%) depending on the concentrations of ions. The removal of Pb(II) by MOL decreased from 97% to 89% after five biosorption/desorption cycles with 0.3?M HCl solution. Freundlich model yielded a better fit for equilibrium data and the pseudo-second-order well described the kinetics of Pb(II) biosorption. FTIR spectra showed that –OH, C–H, –C–O, –C?=?O, and –O–C functional groups were involved in the biosorption of Pb(II). The change in Gibbs free energy (ΔG = ?28.10?kJ/mol) revealed that the biosorption process was favorable and thermodynamically driven. The results suggest MOL as a low cost, environment-friendly alternative biosorbent for the remediation of Pb(II) contaminated water.  相似文献   

19.
In this study we examined the conformation and side chain environments of angiotensins I, II, III, and [Sar1-Ile5-Ala8]angiotensin II using laser Raman spectroscopy. The positions of the amide I bands for all four peptides were found between 1664 and 1673 cm?1. D2O exchange studies confirmed the positions of the amide I and amide III bands. The positions of the amide I bands for all the angiotensins were found at approximately 1665 cm?1 and the amide III bands were all located between 1265 and 1278 cm?1. From the positions and intensities of the amide I and III bands we concluded that all peptides share the same overall conformation consisting of β-turn structure. Spectral analysis indicated that although the spectra for all the peptides were qualitatively identical there was evidence that the angiotensin conformations were more flexible in the aqueous phase than the solid phase. Examination of the 850830 cm?1 tyrosine doublet suggested that the tyrosine residue in the peptides is exposed to the solvent environment and becomes more exposed as the peptide length is decreased. Therefore, there are some localized conformational differences among the angiotensins. The conformational data yielded by this study leads us to conclude that the various biological properties ascribed to the angiotensins are not due to different conformations of the peptides. The biological differences could perhaps be attributed to localized interactions of the individual amino acid residues with themselves and with the hormone receptors.  相似文献   

20.
The heme-based aerotactic transducer (HemAT) is an oxygen-sensor protein consisting of a sensor and a signaling domain in the N- and C-terminal regions, respectively. Time-resolved step-scan FTIR spectroscopy was employed to characterize protein intermediate states obtained by photolysis of the carbon monoxide complexes of sensor-domain, full-length HemAT, and the Y70F (B-helix), L92A (E-helix), T95A (E-helix), and Y133F (G-helix) HemAT mutants. We assign the spectral components to discrete substructures, which originate from a helical structure that is solvated (1638 cm?1) and a native helix that is protected from solvation by interhelix tertiary interactions (1654 cm?1). The full-length protein is characterized by an additional amide I absorbance at 1661 cm?1, which is attributed to disordered structure suggesting that further protein conformational changes occur in the presence of the signaling domain in the full-length protein. The kinetics monitored within the amide I absorbance of the polypeptide backbone in the sensor domain exhibit two distinct relaxation phases (t1 = 24 and t2 = 694 μs), whereas that of the full-length protein exhibits monophasic behavior for all substructures in a time range of t = 1253–2090 μs. These observations can be instrumental in monitoring helix motion and the role of specific mutants in controlling the dynamics in the communication pathway from the sensor to the signaling domain. The kinetics observed for the amide I relaxation for the full-length protein indicate that the discrete substructures within full-length HemAT, unlike those of the sensor domain, relax independently.  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