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 15 毫秒
1.
The far infrared spectra of poly(L -proline) I (190–35 cm?1) and II (400–35 cm?1) were obtained in the solid state at both 300° and 110°K. A significant difference in the region below 100 cm?1 was observed. A very intense band located at 60 cm?1 in the infrared spectrum of form II has no counterpart in form I. This indicates the sensitivity of low-frequency vibrations to the difference in conformation assumed by both forms in the solid state. Additional bands observed in this study are correlated with ir and Raman data previously reported and tentative assignments are made using the results of normal mode calculations (in the single-chain approximation) which have been reported.  相似文献   

2.
M Goodman  N Ueyama  F Naider 《Biopolymers》1975,14(5):901-914
We have studied the nmr spectra of the series of alanine oligopeptides containing a methoxyethoxyethoxyacetyl blocking group on the N-terminal residue and a morpholino blocking group on the C-terminal residue. Spectra were measured in chloroform–trifluoroacetic acid solvent systems. For oligomers with chain lengths of five or more, “double peaks” are observed for the α-CH protons. Addition of trifluoroacetic acid causes the peaks to coalesce. The amount of trifluoroacetic acid necessary for coalescence increases from the pentamer to the nonamer. These findings are general since alanine oligomers with different blocking groups exhibit similar “double peak” phenomena. We explain the “double peak” phenomenon in terms of specific folded forms of the oligopeptides which arise from intramolecular hydrogen bonding. Additional evidence for such hydrogen bonding is presented based upon infrared studies. Slight aggregation probably occurs for the pentamer and hexamer which may stabilize the folded forms.  相似文献   

3.
Fourier-transform infrared (FTIR) spectroscopy was carried out on single colonies of Pediastrum duplex present in air-dried preparations of mixed phytoplankton samples isolated from a eutrophic freshwater lake. FTIR absorption spectra had 12 distinct bands over the wavenumber range 3300–900?cm?1 which were tentatively assigned to a range of chemical groups, including –OH (residual water, wavenumber 3299?cm?1), –CH2 (lipid, 2924), –C=O (cellulose, 1739), amide (protein, 1650 and 1542), >P=O (nucleic acid, 1077) and –C–O (starch, 1151 and 1077). Measurement of band areas identified residual water, protein and starch as the major detectable constituents. Areas of single bands and combined bands of –CH2, –C–O and >P=O species normalized to protein (to correct for differences in specimen hydration and thickness) showed wide variation between colonies, indicating environmental heterogeneity. Correlation analysis demonstrated close statistical associations between different molecular species. Particularly high levels of correlation between bands 3/4 (CH2), 6/7 (amide) and 8/9 (–CH3) was consistent with their joint origin from the same molecular species. The isolation of bands 11 and 12 in the correlation pattern was confirmed by factor analysis, suggesting that variation in the level of starch is statistically unrelated to other macromolecules being monitored. The use of FTIR spectroscopy to characterize an algal micro-population within mixed phytoplankton has potential for future studies on biodiversity and environmental interactions at the species level.  相似文献   

4.
E G Bendit 《Biopolymers》1966,4(5):539-559
A number of basic features of the infrared spectrum of keratin have been confirmed and some new features have been found. In the 3-μ region, the amide A frequency of helical material in α-keratin at 3286 cm.?1 is close to the expected value, but that of the crystalline phase in α-keratin, near 3270 cm.?1, is lower than had previously been reported. The noncrystalline phase absorbs in the vicinity of 3300 cm.?1 or above, and this causes the low-intensity component of the amide A band in both α- and β-keratin to occur at higher frequencies than those of the high-intensity component. In the 6-μ region, the amide II frequency of noncrystalline material is below 1525 cm.?1. Keratin denatured in lithium bromide, after washing out the reagent, appears to have a considerable helix content, possibly as much as that of the original protein. Hydration causes significant spectral changes. In the 6-mu; region, the frequency of the amide I band of crystalline material is lowered, while that of the amide II band is increased, both by a few wavenumbers; the amide II frequency of noncrystaline material is also increased by a few wavenumbers. In the 3-μ region, no significant change is observed in the amide A frequency of crystalline material, while the frequency of the noncrystaline material is reduced. These spectral changes are interpreted in terms of a weak association of water with main-chain carbonyl groups in the crystalline phase, while in the noncrystaline phase it is thought likely that water molecules form hydrogen-bond bridges between polypetide chains. The absorption coefficient of the amide A band and the integrated absorption intensities of the amide A, I, and II bands do not vary appreciably in the three forms of keratin investigated.  相似文献   

5.
G D Fasman 《Biopolymers》1966,4(5):509-519
Poly-O-acetyl-hydroxy-L -proline, forms I and II have been studied by optical rotatory dispersion (ORD) and ultraviolet spectrophotometry in solution and in the solid state. Cotton effects of opposite sign, but not mirror images, were observed in the 250 mμ region for the two forms (Form I, peak 278 mμ; crossover, 254 mμ; trough, 244 mμ: Form II, trough 270 mμ; crossover, 248 mμ; peak, 238 mμ). Thus, the Cotton effects for a right-handed and left-handed helix have been shown to be opposite for the proline type helices I and II. The ORD of films of form I was found to have a positive Cotton effect further into the ultraviolet region with peak at 218 mμ. Absorption spectra showed a shift of 8 mμ in the absorption peak in the 200 mμ region for the two forms (form. I, 211 mμ; form II, 203 mμ). A shoulder was demonstrated in the film absorption spectra in the 250 mμ region where the Cotton effects are found. The mixing of the n, π* and π, π* states of the amide chromophore and n, π* state of the ester chromophore was suggested as being responsible for the Cotton effects in the 250 mμ region.  相似文献   

6.
Synthesis and optical studies of L-methionine oligopeptides in solution   总被引:1,自引:0,他引:1  
F Naider  J M Becker 《Biopolymers》1974,13(5):1011-1022
A series of L -methionine oligomers [BOC-(Met)n-OMe] (n = 2–7) and the corresponding diastereomeric di- and tripeptides were synthesized using the mixed anhydride method. Oligomers prepared in this manner were optically pure and were obtained in reasonable yield. Preliminary optical examination of the peptides suggests that secondary structures may begin forming in the pentamer or hexamer in trifluoroethanol. BOC-(Met)4-OMe and BOC-(Met)5-OMe were also synthesized using an insoluble resin containing BOC-L -methionine as the nitrophenol active ester.  相似文献   

7.
The results of the measurement of the far-ultraviolet absorption spectra of L -proline oligomers in water and acetonitrile are summarized as follows. The monomer has an absorption maximum at 182.5 mμ in acetonitrile. The absorption maximum of the dimer is found at 185 mμ and a shoulder appears around 200 mμ, that is, splitting of the absorption spectrum is observed in the dimer. As the degree of polymerization increases, the position of the shoulder shifts toward the wavelength of the absorption maximum of poly-L -proline II, with an accompanying increase in intensity. We may describe the absorption peak around 203 mμ of poly-L -proline II as identical with the shoulder with an increased intensity. By measurements of optical rotatory dispersion and circular dichroic spectra, it was also confirmed that the appearance of the helical conformation commences at the tetramer. When the number of residues is five or greater, the conformation of the helical structure of poly-L -proline II seems to be completed.  相似文献   

8.
Abstract

Poly(dA-dT) poly(dA-dT) structures in aqueous solutions with high NaCl concentrations and in the presence of Ni2+ ions have been studied with resonance Raman spectroscopy (RRS). In low water activity the effects of added 95 mM NiCl2 in solution stabilize the syn geometry of the purines and reorganize the water distribution via local interactions of Ni-water charged complexes with the adenine N7 position. It is shown that RRS provides good marker bands for a left-handed helix: i) a purine ring breathing mode around 630 cm″?1coupled to the deoxyribose vibration in the syn geometry, ii) a 1300-1340 cm?1 region characterizing local chemical interactions of the Ni2+ ions with the adenien N7 position, iii) lines at about 1483-and 1582 cm?1 correlated to the anti/syn reorientation of the adenine residues on B-Z structure transition, iv) marker bands of the thymidine carbonyl group couplings at 1680-and 1733 cm?1 due to the disposition of the thymidine residues in the Z helix specific geometry. Hence poly(dA-dT) poly(dA-dT) can adopt a Z form in solution. The Z form observed in alternate purine-pyrimidine sequences does not require G-C base pairs.  相似文献   

9.
Polyvinyl alcohol–sodium alginate (PVA–SA) matrix was fabricated and red algae Jania rubens was embedded for removal of lead from aqueous solutions. The Pb(II) uptake rate was rapid primarily at 1 h and equilibrium was achieved within 2 h. The optimum pH was 5, the data were well fitted by Langmuir and Freundlich models, and RL values are in the range of 0.1–0.38. The sorption capacity (qe) of PVA–calcium alginate (CA)–J. rubens matrix increased from 10.77 to 37.195 mg g?1 with increasing Pb(II) concentration from 24.86 to 98.75 mg L?1 at the temperature of 30°C and pH 5. The sorption capacity (qe) and maximum biosorption (qm) were noted as 37.179 ± 0.32 and 71.43 mg/g, respectively. The adsorption process was well described by pseudo-second-order model. The reaction is endothermic, is spontaneous, and increases in randomness. The functional groups present on matrix, i.e., –OH, –C–N, –C–O,–CO–NH, –NH2, –SH, and –C–OH, were intensely involved in the process. Scanning electron microscopy results revealed the morphological changes due to adsorption of Pb(II) on and inside of PVA–CA–J. rubens matrix. Desorption study indicates the efficient regeneration of PVA–CA–J. rubens biomass matrix for three cycles and is a promising matrix for removal of Pb(II) and can be used in continuous systems.  相似文献   

10.
Far-infrared spectra of poly-L -alanines having the α-helical conformation and the β-form structure were measured. The spectra of glycine–L -alanine copolymer, silk fibroin, and copoly-D ,L -alanines with different D :L compositions were also measured. In addition to the bands so far reported, four bands at 190, 107, 120, and 90 cm?1were found for the α-helix conformation and the two bands at 442 and 247 cm?1 were found for the β form. The 442 cm?1 band consists of the parallel 432 cm?1 and perpendicular 445 cm?1 bands. The 247 cm?1 band is well defined and has strong dichroism parallel to the direction of stretching. These two bands appear also for silk fibroin and glycine–L -alanine copolymer. All the far-infrared bands of copoly-D ,L -alanines can be interpreted as α-helix bands, the three peaks at 580, 478, and 420 cm?1 being ascribed to the D -residue incorporated into the right-handed α-helix or to the L -residue in the left-handed α-helix.  相似文献   

11.
Staphylococcus saprophyticus strains ATCC 15305, ATCC 35552, and ATCC 49907 were found to require l-proline but not l-arginine for growth in a defined culture medium. All three strains could utilize l-ornithine as a proline source and contained l-ornithine aminotransferase and Δ1-pyrroline-5-carboxylate reductase activities; strains ATCC 35552 and ATCC 49907 could use l-arginine as a proline source and had l-arginase activity. The proline requirement also could be met by l-prolinamide, l-proline methyl ester, and the dipeptides l-alanyl-l-proline and l-leucyl-l-proline. The bacteria exhibited l-proline degradative activity as measured by the formation of Δ1-pyrroline-5-carboxylate. The specific activity of proline degradation was not affected by addition of l-proline or NaCl but was highest in strain ATCC 49907 after growth in Mueller–Hinton broth. A membrane fraction from this strain had l-proline dehydrogenase activity as detected both by reaction of Δ1-pyrroline-5-carboxylate with 2-aminobenzaldehyde (0.79 nmol min−1 mg−1) and by the proline-dependent reduction of p-iodonitrotetrazolium (20.1 nmol min−1 mg−1). A soluble fraction from this strain had Δ1-pyrroline-5-carboxylate dehydrogenase activity (88.8 nmol min−1 mg−1) as determined by the NAD+-dependent oxidation of dl1-pyrroline-5-carboxylate. Addition of l-proline to several culture media did not increase the growth rate or final yield of bacteria but did stimulate growth during osmotic stress. When grown with l-ornithine as the proline source, S. saprophyticus was most susceptible to the proline analogues L-azetidine-2-carboylate, 3,4-dehydro-dl-proline, dl-thiazolidine-2-carboxylate, and l-thiazolidine-4-carboxylate. These results indicate that proline uptake and metabolism may be a potential target of antimicrobial therapy for this organism.  相似文献   

12.
Abstract

The interaction of DNA and RNA with Cu(II), Mg(II), [Co(NH3)6]3+ [Co(NH3)5Cl]2+ chlorides and, cis- and trans-Pt(NH3)2Cl2 (CIS-DDP, trans-DDP) has been studied by Fourier Transform Infrared (FT-IR) spectroscopy and a correlation between metal-base binding and conformational transitions in the sugar pucker has been established. It has been found that RNA did not change from A-form on complexation with metals, whereas DNA exhibited a B to Z transition. The marker bands for the A-form (C′3-endo-anti conformation) were found to be near 810–816 cm?1, while the bands at 825 and 690 cm?1 are marker bands for the B- conformation (C′2-endo, anti), The B to Z (C3-endo, syn conformation) transition is characterized by the shift of the band at 825 cm?1 to 810–816 cm?1 and the shift of the guanine band at 690 cm?1 to about 600–624 cm?1.  相似文献   

13.
E G Bendit 《Biopolymers》1966,4(5):561-577
A number of new bands have been found in the spectra of deuterated α- and β-keratin. In particular, the deuteration difference spectrum has been useful for the determination of frequencies of previously unsuspected bands. Thus it is found that the amide A and II frequencies of the nonhelical component in α-keratin occur at 3310–3320 and 1520 cm.?1, respectively, and that both bands exhibit dichroism consistent with polypeptide chains which have a measure of alignment parallel to the fiber axis. The parallel dichroism of the amide II′ band of this phase at about 1435 cm.?l also indicates some alignment. A nondichroic residual band at 1513 cm.?1 in highly deuterated α-keratin is assigned to the tyrosine residue, as a sharp band near this frequency is found in the spectrum of polytyrosine. The ν‖(o) component of the α-helix is weak or absent in α-keratin, and the relatively sharp band observed near this frequency is thought to be due to the tyrosine residue, while its dichroism is caused by the presence of dichroic nonhelical material. A band near 1575 cm.?1 in deuterated α- and β-keratin is tentatively assigned to the deuterated guanidinium group of arginine. This band becomes progressively more prominent during deuteration, which indicates that some arginine side chains arc slow to exchange, possibly because their environment prevents interaction with D2O. The deuteration difference spectrum also shows that, contrary to earlier views, helical material in α-keratin exchanges significantly during the early stages of deuteration, although at a slower rate than the nonhelical material, while part of the nonhelical phase does not exchange as rapidly as had been thought and makes a contribution even after many hours or days.  相似文献   

14.
In order to clarify a role of the proline residue at near cysteine and histidine positions of plastocyanin and azurin, N-mercaptoacetylglycyl-L-prolyl-L-histidine has been synthesized as an analogous ligand of blue copper sites and the spectroscopic properties of its Cu(II) complex compared with those of the N-mercaptoacetylglycylglycyl-L-histidine-Cu(II) complex. In the present tetrapeptide-Cu(II) complexes, the exchange of the glycine of the third position by the proline residue effects a red shift(80 nm) of the visible absorption and a decrease (192→75×10?4cm?1) of the copper hyperfine splitting. The introduction of proline residue induces a change of the complex geometry from D4h to Td symmetries.  相似文献   

15.
Infrared absorption spectra for a number of polysaccharides and their nitrated derivatives have been obtained. The frequency range 730–960 cm?1 is useful for identification of the polysaccharides, and the region 900–1350 cm?1 is more suitable for distinguishing the nitrated materials. The strong intensity of the nitrate bands limits the interpretation of spectra below 960 cm?1, but above this frequency the absorption bands of nitrated polysaccharides are generally sharper and more clearly defined than the corresponding bands of the parent polysaccharides. Data on the COC bridge, CC ring, CO, and COH frequencies and on the CH deformation and stretching frequencies have been obtained. The use of i.r. spectroscopy for the quantitative determination of nitrate groups in nitrated polysaccharides is discussed.  相似文献   

16.
For infrared absorption measurements, the following five isotopic polyglycines have been prepared: ordinary polyglycine (—NHCH2CO—)n, N-deuterated polyglycine (—NDCH2CO—)n, C-deuterated polyglycine (—NHCD2CO—)n, completely deuterated polyglycine (—NDCD2CO—)n, and N15-substituted polyglycine (—15NHCH2CO—)n. Infrared spectra have been observed both in the I and II forms of each of these five isotopic polyglycines in the spectral region of 4000–300 cm.?1. On the basis of the comparison of these spectra with each other, a nearly complete set of assignments of the observed bands of polyglycines has been given.  相似文献   

17.
The relationship between published vicinal proton–proton coupling constants and the pseudorotation properties of the pyrrolidine ring in L -proline, 4-hydroxy-L -proline, 4-fluoro-L -proline, and several linear and cyclic model proline peptides is investigated. Compared to earlier studies, several important improvements are incorporated: (1) a new empirical generalization of the classical Karplus equation is utilized, which allows a valid correction for the effects of electronegativity and orientation of substitutents on 3JHH; (2) an empirical correlation between proton–proton torsion angles and the pseudorotational parameters P and τm is derived; and (3) the best fit of the conformational parameters to the experimental coupling constants is obtained by means of a computerized iterative least-squares procedure. Two pseudorotation ranges were considered, classified as type N (χ2 positive sign) and type S (χ2 negative sign). The conformational equilibrium is fully described in terms of four geometrical parameters (PN, τN, PS, τS) and the equilibrium constant K. The present results indicate that, in general, the geometrical properties found in x-ray studies of proline and hydroxyproline residues are well preserved in solution. Several novel features are encountered, however. It is demonstrated that the proline ring occurs in a practically 1:1 conformational equilibrium between well-defined N- and S-type forms. Introduction of an amide group at the C-terminal end has no observable effect on this equilibrium, but the formation of a peptide bond at the imino nitrogen site results in a pronounced, but not exclusive, preference for an S-type form which is roughly 1.1 kcal/mol more stable than its N-type counterpart. The hydroxyproline ring system in neutral or acidic medium displays a pure N-type state, but N-acetylation results in the appearance of a minor (S-type) conformation. Cyclic proline dipeptides similarly exist in a biased conformational equilibrium. The major form (77–88%) corresponds to the N-type conformer observed in the solid state; the minor S-form has not been observed before. In contrast, cyclic hydroxyproline dipeptides display complete conformational purity. Ranges of endocyclic torsion angles deduced for the various classes of pyrrolidine derivatives in solution are presented. Each torsion appears confined to a surprisingly narrow range, comprising about 4°–8° in most cases. In all, the proline ring is far less “floppy” than hitherto assumed.  相似文献   

18.
The infrared absorption spectra of native (potato) and new modified starches were observed in about 4500 cm?1~670 cm?1 region using film technique as a new attempt. Deuteration was attempted to investigate the physicochemical properties and to assign the absorption bands partially. And the absorption band attributed to the H2O molecule was explained. The absorption band, which appeared in the modified starch only and which was thought as one great difference between two starches seems to be related to CO.  相似文献   

19.
H. N. Cheng  F. A. Bovey 《Biopolymers》1977,16(7):1465-1472
By means of carbon-13 nmr (at 25 MHz) the trans/cis conformer ratio in glycyl-L -proline has been measured in aqueous (D2O) solution over the temperature range 33–96°C. It is found that ΔH0 = ?4.2 kJ/mole and ΔS0 = ?9.7 J/mole/K. Measurements of the T1 values for the proline ring carbons yielded values consistent with a fast puckering process involving both the β- and γ-carbons. Measurements of the rate of cis-trans conformational interconversion in glycyl-L -proline, using complete line-shape analysis for the glycyl α-carbon resonance, gave values for the transcis isomerization as follows: ΔH = 83.5 ± 0.2 kJ/mole; ΔS = 0.0 ± 10 J/mole/K. A more approximate determination from coalescence temperature observations gave a value of ΔG of 82.0 ± 0.4 kJ/mole for this process in acetyl-L -proline in aqueous solution. The presence of 12M NaSCN lowered this barrier by ca. 2.6 kJ/mole. Such measurements are relevant to present theoretical models of the denaturation-renaturation processes in proteins, in which proline residues may play a key role.  相似文献   

20.
We present a study of soil organic carbon (SOC) inventories and δ13C values for 625 soil cores collected from well‐drained, coarse‐textured soils in eight areas along a 1000 km moisture gradient from Southern Botswana, north into southern Zambia. The spatial distribution of trees and grass in the desert, savannah and woodland ecosystems along the transect control large systematic local variations in both SOC inventories and δ13C values. A stratified sampling approach was used to smooth this variability and obtain robust weighted‐mean estimates for both parameters. Weighted SOC inventories in the 0–5 cm interval of the soils range from 7 mg cm?2 in the driest area (mean annual precipitation, MAP=225 mm) to 41±12 mg cm?2 in the wettest area (MAP=910 mm). For the 0–30 cm interval, the inventories are 37.8 mg cm?2 for the driest region and 157±33 mg cm?2 for the wettest region. SOC inventories at intermediate sites increase as MAP increases to approximately 400–500 mm, but remain approximately constant thereafter. This plateau may be the result of feedbacks between MAP, fuel load and fire frequency. Weighted δ 13C values decrease linearly in both the 0–5 and 0–30 cm depth intervals as MAP increases. A value of –17.5±1.0‰ characterizes the driest areas, while a value of ?25±0.7‰ characterizes the wettest area. The decrease in δ 13C value with increasing MAP reflects an increasing dominance of C3 vegetation as MAP increases. SOC in the deeper soil (5–30 cm depth) is, on average, 0.4±0.3‰ enriched in 13C relative to SOC in the 0–5 cm interval.  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