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.
In order to determine the ionization state of the 5'-phosphate of bound pyridoxal phosphate, a Fourier transform infrared spectroscopic study of cytosolic aspartate aminotransferase has been carried out. Dianionic and monoanionic phosphate monoesters give rise to two bands each in the infrared spectrum [Shimanouchi, T., Tsuboi, M., & Kyogoku, Y. (1964) Adv. Chem. Phys. 8, 435-498]. These bands can be identified in infrared spectra of the free coenzyme in solution. Due to interfering bands arising from the protein, only the band assigned to the symmetric stretching of the dianionic phosphate is observed in holoenzyme solutions. The integrated intensity of this band does not change with pH in the range 5.3-8.6, while for free pyridoxal phosphate, the integrated intensity of the same band changes with pH according to the pK value expected for the 5'-phosphate group in solution. Moreover, the value of the integrated intensity for the bound cofactor is close to the value given by free cofactor at pH 8-9. These results suggest that the 5'-phosphate of the bound cofactor remains mostly dianionic throughout the investigated pH range and disfavor other interpretations in terms of ionization of the phosphate group on the basis of the nuclear magnetic resonance 31P chemical shift-pH titration curve of holoenzyme [Schnackerz, K. D. (1984) in Chemical and Biological Aspects of Vitamin B6 Catalysis (Evangelopoulos, E. A., Ed.) Part A, pp 195-208, Alan R. Liss, New York].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
Phosvitin, a highly phosphorylated glycoprotein, represents the major fraction of hen egg yolk phosphoproteins. Circular dichroism, Fourier transform infrared spectroscopy, and Fourier transform infrared photoacoustic and fluorescence spectroscopic methods were employed to determine the secondary structure of the protein in both the solid and solution phases. This was supplemented by a Chou-Fasman type of predictive algorithm for the first 25 residues at the N terminus of the dephosphorylated protein. A three-compartment model consisting of alpha-helical, beta-sheet, and beta-turn components with beta-turns occurring at the interface between alpha-helical and beta-sheet regions in the proximity of O-phosphoserine residues is suggested from the combined analyses. Beta-sheets appear to be the dominant secondary structural component in phosvitin in the solid and solution phases. The suggested model bears many similarities to other phosphoproteins reported in the literature. The secondary structure of phosvitin is observed to be sensitive to environmental factors as previously reported although the present studies differ in some respects from earlier results. Preliminary results suggest that Ca2+ ions trigger a decrease in beta-sheet structure at pH 2.  相似文献   

3.
By measuring the rhodopsin--bathorhodopsin, isorhodopsin--bathorhodopsin, rhodopsin--isorhodopsin and rhodopsin--meta-II difference spectra with the method of Fourier-transform infrared spectroscopy we have identified the C = N stretching vibration of the protonated retinylidene Schiff base of rhodopsin, isorhodopsin and bathorhodopsin. In contrast to resonance Raman spectroscopy additional strong bands were observed between 1700 cm-1 and 1620 cm-1. Most of them depend on the isomeric state of the chromophore. The origin of these bands will be discussed. In the fingerprint region isorhodopsin and bathorhodopsin are quite similar but no similarities with infrared spectra of model compounds of any isomeric composition are observed. Therefore, no conclusions on the isomeric state of the retinal in bathorhodopsin can be drawn. We provide evidence for the modification of one or two carboxylic group(s) during the rhodopsin--bathorhodopsin and isorhodopsin--bathorhodopsin transition.  相似文献   

4.
The effects of alcohols (methanol, ethanol, and n-butanol) on the hydrogen bonding of dipalmitoylphosphatidylcholine (DPPC) were studied by Fourier-transform infrared spectroscopy (FTIR) in water-in-oil (carbon tetrachloride) reversed micelles. The bound O-H stretching mode of water, bonded to DPPC, appeared as a broad band at around 3400 cm-1. The O-H bending mode of this complex appeared as a weak broad band at 1644 cm-1. No free O-H signal was observed. When alcohols were added, a part of DPPC-bound water was replaced by the alcohols. The released 'free' water appeared at 3680 cm-1. This free O-H stretching band represents water-alcohol complex. A new broad band of O-H stretching appeared at 3235 cm-1, which represents the alcohol molecules bound to the phosphate moiety of DPPC. When the alcohol concentration was increased, the intensities of the free O-H stretching and bending bands increased. The P = O- antisymmetric stretching band at 1238 cm-1 became broader and shifted to lower frequencies. This means that alcohols interacted with the phosphate moiety and replaced the bound water. In the deconvoluted spectra of the C = O stretching mode, the ratio between the free sn-2 and the hydrogen-bonded sn-2 bands increased; a part of the bound water at the sn-2 carbon in the glycerol skeleton is also released and the free sn-2 signal increased. From the change in the intensity of the P = O- stretching band, the partition coefficients of alcohols between the phosphate region of DPPC and water were estimated: methanol 7.8, ethanol 16.7 at 22.0 degrees C in mole fraction bases. In molality, these values translates into methanol 0.21 and ethanol 0.45. These results indicate that short-chain alcohols interact with lipid membranes at the phosphate moiety at the hydrophilic head, weaken the membrane-water interaction, and destabilize membranes.  相似文献   

5.
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   

6.
Ser-473 is solely phosphorylated in vivo in the tail region of neurofilament L (NF-L). With peptides including the native phosphorylation site, it was not possible to locate responsible kinases. We therefore adopted full-length dephosphorylated NF-L as the substrate, and employed MALDI/TOF (matrix-assisted laser desorption and ionization/time of flight) mass spectrometry and a site-specific phosphorylation-dependent antibody recognizing Ser-473 phosphorylation. The antibody showed that casein kinase I (CK I) as well as casein kinase II (CK II) phosphorylated Ser-473 in vitro, while neither GSK-3beta nor calcium/calmodulin-dependent protein kinase II did so. However, the mass spectra of the tail fragments of the phosphorylated NF-L indicated that CK II was the kinase mediating Ser-473 phosphorylation in vitro as opposed to CK I, because CK I phosphorylated another site as well as Ser-473 in vitro. The antibody also demonstrated that NF-L phosphorylated at Ser-473 was abundant in the neuronal perikarya of the rat cortex, indicating that phosphorylation of Ser-473 may take place there. This result may support the suggestion that CK II is the kinase responsible for Ser-473 phosphorylation. Despite many reports showing that CK I mediates phosphorylation of neurofilaments, CK II may phosphorylate NF-L in vivo.  相似文献   

7.
(1) High-resolution 31P-NMR was used to study the environment of the phosphoserine residues of the phosphoproteins, alpha s1-casein B, beta-casein A2 and beta-casein C. For reference purposes 31P-NMR spectra of phosvitin and ovalbumin were also collected. (2) 31P resonances were assigned to specific phosphoserine residues as a result of comparisons of the high-resolution 31P-NMR spectra for alpha s1- and beta-caseins and for peptide fragments of these proteins obtained by cyanogen bromide and trypsin cleavage. (3) Measurements of the enhancement of the relaxation rate for water protons (1H) on addition of Mn2+ to alpha s1-casein B and to a fragment alpha s1-CN3, obtained by cyanogen bromide cleavage, gave approximate pK values for the binding groups and suggest the possibility of a conformational change induced by varying the concentration of divalent cation.  相似文献   

8.
The region of the infrared spectrum corresponding to C-2H stretching vibrations (2050-2250 cm-1) has been examined for liposomes composed of dimyristoylphosphatidylcholine deuterated specifically at the methyl ends of either one (sn-2) or both the fatty acyl chains. This label is intended to provide information on lipid dynamics in the contact region between monolayers. The two most prominent bands observed correspond, respectively, to antisymmetric (2212 cm-1) and symmetric (2075 cm-1) C-2H stretching vibration. The antisymmetric band consists of two overlapping peaks, whose positions vary with the gel or liquid-crystalline state of the lipid. The separation between the peaks making up the antisymmetric band increases with temperature, and is maximum above the Tc transition temperature; this rules out the previously proposed assignment of these two peaks to different rotational modes of the methyl group relative to the adjacent methylene. The position and width of the symmetric band at 2075 cm-1 are also sensitive to the physical state of the lipid. The presence of cholesterol at an equimolar ratio with the phospholipid abolishes all the phase-dependent changes observed. The intrinsic polypeptide gramicidin A, at a 5:1 lipid/peptide mol ratio, is seen to enlarge the lipid thermotropic transition, with small effects above Tc. Cytochrome c, an extrinsic protein, at a 10:1 mole ratio, does not modify the phase-dependent behaviour of the terminal methyl groups, but consistently shifts all the observed bands to lower-frequency positions, which suggests a long-range effect of the protein along the phospholipid fatty acyl chains.  相似文献   

9.
The ionization state of the phosphate group bound at the aspartate aminotransferase apoenzyme's active site has been investigated utilizing Fourier-transform infrared spectroscopy following the band corresponding to the symmetric stretching of the dianionic phosphate. Unlike free phosphate, when inorganic phosphate is bound at the enzyme's active site, the integrated intensity value of the dianionic band does not change with pH within the studied range, and this value is similar to that for free dianionic phosphate at pH 8.3. From these results, we propose a dianionic state for the phosphate ion bound to cytosolic aspartate aminotransferase throughout the pH range of 5.7-8.3. The presence of other anions such as acetate and chloride or the substrate aspartate and its analogues produces a pH-dependent phosphate removal from the active site which is favored at low pH values. Elimination of the charged primary amine at the active-site Lys-258, through formation of a Schiff base with pyridoxal or chemical modification by carbamylation, also produces a pH-independent phosphate release. These results are interpreted as Lys-258 together with the active-site alpha-helix and other residues may be involved in stabilizing phosphate as a dianion in the apoenzyme phosphate pocket which anchors the phosphate ester of pyridoxal phosphate in the holoenzyme. It is proposed that the dianionic phosphate contributes to the apoenzyme's thermal stability through formation of strong hydrogen bond and salt bridges with the amino acid residues forming the phosphate binding pocket with assistance of Lys-258, and other active-site cationic components.  相似文献   

10.
Salmonella enterica serovars are prevalent foodborne pathogens responsible for high numbers of salmonellosis each year. Complex Fourier-transform infrared (FTIR) spectra offer unique biochemical fingerprints of bacteria with bands due to major cellular components. Growth media effects on discrimination of Salmonella serovars by FTIR spectroscopy were investigated and a novel sample preparation technique was developed. S. enterica strains from six serovars were grown on xylose lysine desoxycholate (XLD), Miller-Mallinson (MM), and plate count (PCA) agar as a control (37 degrees C, 24 h). Isolated colonies were suspended in 50% acetonitrile and centrifuged; the remaining pellet was placed on an AMTIR (attenuated total reflectance) crystal and dried under vacuum. Classification models (Soft Independent Modeling of Class Analogy, SIMCA), generated from derivatized infrared spectra (1300-900 cm-1 or 1200-900 cm-1), successfully discriminated among Salmonella strains with major discrimination from 1000-970 cm-1 associated to stretching modes of O-specific polysaccharide chains of lipopolysaccharides. Sample treatment with acetonitrile enhanced safe handling of the bacteria, removed interfering signals and improved the discriminating ability of SIMCA. All media were able to discriminate the S. enterica strains studied, varying in discriminating peaks and class distances in SIMCA classification. This methodology, with the production of large libraries of pathogenic bacteria, could be applied for the rapid monitoring of bacterial contamination in food with minimal sample manipulation.  相似文献   

11.
Two protein kinases active on casein and phosvitin were partially purified from the soluble fraction of ejaculated bovine spermatozoa. They were operationally termed casein kinase A and B based on the order of their elution from a phosphocellulose column. CK-A showed an approximate molecular mass of 38 kDa, and it phosphorylated serine residues of casein and phosvitin utilizing ATP as a phosphate donor (Km 19 microM). Enzyme activity was maximal in the presence of 10 mM MgCl2, whereas it decreased in the presence of spermine, polylysine, quercetin, and NaCl (20-250 mM). CK-B seemed to have a monomeric structure of about 41 kDa; it underwent autophosphorylation and cross-reacted with polyclonal antibodies raised against recombinant alpha, but not beta, subunit of human type 2 casein kinase. It phosphorylated both serine and threonine residues of casein and phosvitin, utilizing ATP (Km 12 microM) but not GTP as a phosphate donor. Threonine was more affected in the phosphorylated phosvitin than in the partially dephosphorylated substrate. CK-B was active toward the synthetic peptide Ser-(Glu)5 and calmodulin (in the latter case, in the presence of polylysine), and it was activated by spermine, polylysine, MgCl2 (30 mM), and NaCl (20-400 mM). The activity of the enzymes was not affected by cAMP, or the heat-stable inhibitor of the cAMP-dependent protein kinase, or calcium.  相似文献   

12.
Fourier-transform infrared spectroscopy is a valuable method for the study of protein conformation in solution primarily because of the sensitivity to conformation of the amide I band (1700-1620 cm-1) which arises from the backbone C = O stretching vibration. Combined with resolution-enhancement techniques such as derivative spectroscopy and self-deconvolution, plus the application of iterative curve-fitting techniques, this method provides a wealth of information concerning protein secondary structure. Further extraction of conformational information from the amide I band is dependent upon discerning the correlations between specific conformational types and component bands in the amide I region. In this paper, we report spectra-structure correlations derived from conformational perturbations in bovine trypsin which arise from autolytic processing, zymogen activation, and active-site inhibition. IR spectra were collected for the single-chain (beta-trypsin) and once-cleaved, double-chain (alpha-trypsin) forms as well as at various times during the course of autolysis and also for zymogen, trypsinogen, and beta-trypsin inhibited with diisopropyl fluorophosphate. Spectral differences among the various molecular forms were interpreted in light of previous biochemical studies of autolysis and the known three-dimensional structures of the zymogen, the active enzyme, and the DIP-inhibited form. Our spectroscopic results from these proteins in D2O imply that certain loop structures may absorb in the region of 1655 cm-1. Previously, amide I' infrared bands near 1655 cm-1 have been interpreted as arising solely from alpha-helices. These new data suggest caution in interpreting this band. We have also proposed that regions of protein molecules which are known from crystallographic experiments to be disordered absorb in the 1645 cm-1 region and that type II beta-turns absorb in the region of 1672-1685 cm-1. Our results also corroborate assignment of the low-frequency component of extended strands to bands below 1636 cm-1. Additionally, the results of multiple measurements have allowed us to estimate the variability present in component band areas calculated by curve fitting the resolution-enhanced IR spectra. We estimate that this approach to data analysis and interpretation is sensitive to changes of 0.01 unit or less in the relative integrated intensities of component bands in spectra whose peaks are well resolved.  相似文献   

13.
I.r. absorption bands associated with the functional groups of carboxylic acid derivatives are useful for the analysis of alginates and pectins. The ester, amide, and uronate contents of pectins and the uronate content of alginates were determined, respectively, from the ester-carbonyl stretching band (1740 cm- minus 1), the amide I band (1650 cm- minus 1), and the carboxylate antisymmetric stretching band (1607 cm- minus 1) obtained from the spectra of solutions in D2O-phosphate buffer. The results are accurate to within plus or minus 2-4%, are self consistent, and agree well with the few reliable results that are available. The method should be applicable for the determination of carboxylic acid derivatives in other polysaccharides.  相似文献   

14.
Interactions between the nicotinic acetylcholine receptor (nAChR) and phosphatidic acid (PA) are bidirectional in that membranes containing PA are effective at stabilizing an agonist-responsive nAChR, whereas incorporation of the nAChR into the same membranes leads to a substantial increase in lipid lateral packing density. A previous study suggested that the ability of PA to adopt a dianionic ionization state is key. We monitored the ionization state of PA in both reconstituted and protein-free membranes. In model membranes composed of PA and 3:2 (mol/mol) phosphatidylcholine (PC)/PA, the monoanionic-to-dianionic transition of PA was detected with a pKa of 8.7 and 6.5, respectively. In the reconstituted 3:2 PC/PA membranes, however, PA was stabilized in a monoanionic state at pH values up to 10. Although dianionic PA does not play a role in nAChR function, we found that both the stabilization of monoanionic PA and the concentration of other cations at the bilayer surface can account for changes in bilayer physical properties that are observed upon incorporation of the nAChR into 3:2 PC/PA membranes. A nAChR-induced concentration of cations at the bilayer surface likely mediates interactions between the nAChR and the anionic lipids in its membrane environment.  相似文献   

15.
We have carried out a Fourier transform infrared spectroscopic study of mitochondrial aspartate aminotransferase in the spectral region where phosphate monoesters give rise to absorption. Infrared spectra in the above-mentioned region are dominated by protein absorption. Yet, below 1020 cm-1 protein interferences are minor, permitting the detection of the band arising from the symmetric stretching of dianionic phosphate monoesters [T. Shimanouchi, M. Tsuboi, and Y. Kyogoku (1964) Adv. Chem. Phys. 8, 435-498]. The integrated intensity of this band in several enzyme forms (pyridoxal phosphate, pyridoxamine phosphate, and sodium borohydride-reduced, pyridoxyl phosphate form) does not change with pH in the range 5-9. This behavior contrasts that of free pyridoxal phosphate (PLP) and pyridoxamine phosphate (PMP) in solution, where the dependence of the same infrared band intensity with pH can be correlated to the known pK values for the 5'-phosphate ester in solution. The integrated intensity value of this infrared band for the PLP enzyme form before and after reduction with sodium borohydride is close to that given by free PLP at pH 8-9. These results are taken as evidence that in the active site of mitochondrial aspartate aminotransferase the 5'-phosphate group of PLP remains mostly dianionic even at a pH near 5. Thus, it is suggested that the chemical shift changes associated with pH titrations of various PLP forms reported in a previous 31P NMR study of this enzyme [M. E. Mattingly, J. R. Mattingly, and M. Martinez-Carrion (1982) J. Biol. Chem. 257, 8872] are due to the fact that the phosphorus chemical shift senses the O-P-O bond distortions induced by the ionization of a nearby residue. Since no chemical shift changes were observed in pH titrations of the PMP forms (lacking an ionizable internal aldimine) of this isozyme, the Schiff base between PLP and Lys-258 at the active site is the most likely candidate for the ionizing group influencing the phosphorus chemical shift in this enzyme.  相似文献   

16.
The 1000-1300 cm-1 region of the infrared spectrum of dipalmitoylphosphatidylcholine (DPPC) and other phosphate-containing molecules has been studied by the Fourier-transform technique. Three absorption bands have been assigned to various vibrational modes of the DPPC phosphate group, with maximum wavenumbers at 1060, 1086 and 1222 cm-1. These values are the same above and below Tc of the phospholipid. Dehydration produces band-shifts toward higher wavenumbers .  相似文献   

17.
Two different phosphopeptide (PPP) fragments derived from partially dephosphorylated hen egg yolk phosvitin were prepared by tryptic digestion, and their Ca2+ binding property compared with that of commercial casein phosphopeptides (CPP). The smaller fragment of less than 1 kDa and O-phospho-1-serine did not bind Ca2+ to any significant extend, while PPP of 1-3 kDa showed a higher ability than CPP to render soluble calcium. The results show that not only the phosphoserine residues are critical for Ca2+ binding, but also the molecular size of the phosphopeptides.  相似文献   

18.
Substrate specificity of Gaucher spleen phosphoprotein phosphatase   总被引:1,自引:0,他引:1  
The spleen in Gaucher's disease contains elevated levels of two distinct acid phosphatases. One of the isoenzymes, a tartrate-resistant type 5 acid phosphatase which we have designated SPII acid phosphatase, possesses considerable phosphoprotein phosphatase activity. The enzyme dephosphorylates phosvitin and casein at specific rates (V) of 38.6 and 45.0 units/mg, respectively. The dephosphorylation of the oligophosphoproteins as well as various fragments of phosvitin, histories, and monophosphopeptides was studied kinetically. Positive cooperativity (Hill coefficient = 1.3–2.0) was observed for the dephosphorylation of phosvitin and casein as well as for the dephosphorylation of fragments of phosvitin which contained as few as two vicinal phosphoserine residues. In contrast, the hydrolysis of phosphomonoesters such as o-phosphorylserine or various monophosphopeptides exhibited typical Michaelis-Menten kinetics. Cooperativity appears to depend upon the substrate rather than the enzyme. The cooperativity of dephosphorylation was not affected by altering the secondary structure of phosvitin from a random to β conformation or by acetylation of the protein; however, acetylated phosvitin was dephosphorylated more rapidly (V = 50.8 units/mg) than native phosvitin indicating that the very basic phosphatase enzyme (pI = 8.5) prefers more acidic phosphoproteins as substrates rather than basic proteins such as histone (V= 0.0013 unit/mg). A monophosphohexa-peptide (V = 0.47 unit/mg) and monophosphoheptapeptide (V = 0.18 unit/mg) proved to be much poorer substrates than phosvitin, and monophosphoproteins such as glycogen phosphorylase, phosphorylase kinase, and glycogen synthase were not dephosphorylated by the enzyme. Although the phosphatase is active on monophosphopeptides and the presence of flanking amino acids considerably decreases the Km of the enzyme for the phosphoserine residue (up to 100-fold), the enzyme appears to prefer peptide or protein substrates that contain two or more phosphoserine residues in close proximity. Finally, previous results showing the spleen phosphatase to be composed of 16,000- and 20,000-dalton subunits were apparently due to proteolysis during isolation since when 1.0 mm phenylmethylsulfonyl fluoride was included in the isolation media, the enzyme appeared as a single 35,000-dalton species when subjected to polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulfate.  相似文献   

19.
M Tsubaki  T Mogi  H Hori 《FEBS letters》1999,449(2-3):191-195
Azide-binding to the heme-copper binuclear center of bo-type ubiquinol oxidase from Escherichia coli was investigated with Fourier-transform infrared spectroscopy. Deconvolution analyses of infrared spectra of the azide (14N3)-inhibited air-oxidized form showed a major infrared azide antisymmetric stretching band at 2041 cm(-1). An additional band developed at 2062.5 cm(-1) during a longer incubation. Isotope substitutions with terminally 15N-labelled azides did not show a splitting of the major band, indicating that the geometry of the bound azide is mainly in a bridging configuration between high-spin heme o and CuB. The band at 2062.5 cm(-1) showed clear splittings upon substitution with the terminally 15N-labelled azides, indicating the Cu(2+)B-N=N=N structure. Partial reduction of the oxidase with beta-NADH in the presence of azide caused an appearance of new infrared bands at 2038.5 (major) and 2009 (minor) cm(-1). The former band also showed clear splittings in the presence of the terminally 15N-labelled azides, indicating that reduction of low-spin heme b alters the structure of the binuclear center leading to the Fe(3+)o-N=N=N configuration.  相似文献   

20.
In the IR spectra, the coupling of vibrations leads to band splitting and/or bands shifting in opposite directions which provides information on the mutual orientation of groupings. From such band shifts in the range 1800 to 1500 cm-1 one can draw conclusions on the double helix formation of polynucleotides. These band shifts are caused either by vibrational coupling of stretching vibrations within pairs of base residues or by coupling of stretching vibrations with the bending (scissor) vibration of the -NH2 groups; the latter is indicated by band shifts after deuterium substitution within the amino groups. Couplings of phosphate and 1 ibose vibrations in the range 1300 to 1000 cm-1 provide information on the secondary structure of the backbone. In order to obtain information of the structure of the RNA backbone, the IR spectra of poly(ribonucleotides) were studied in neutral media in which they were single-stranded. The shift due to coupling of the band of the 2'OD bending vibration and that of the antisymmetric stretching vibration of the ether group of the ribose residue proves that ribose residues of the backbone are cross-linked via hydrogen bonds. These are formed between the 2'OD or 2'OH groups, respectively, and the O atoms of the ether group of the neighboring ribose residues. This is the reason for the difference between DNA and RNA as regards the 2'OH group. The structure formation caused by these hydrogen bonds results in a stiffening of the RNA backbone. The tendency to form these hydrogen bonds increases in the order poly (U), poly(C), poly (A). This order of secondary structure stabilization is due to an interplay between the influences of (1) the 2'OH hydrogen bonds and (2) the base residues' stacking. Furthermore, the coupling of the antisymmetric stretching vibration of the greater than PO2- groups with a vibration involving the 2'OH group can result in a doublet structure of the band at about 1240 cm-1 if cations with strong fields are present. This probably shows that these cations can turn the greater than PO2-groups-which are usually turned outward at the backbone, as shown by construction of molecular models- towards the basic residues. Thus they cause stiff monohelices which are right-handed screws.  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