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 46 毫秒
1.
The binding of vanadyl to a porcine and bovine testicular S-100-like protein and to calmodulin was demonstrated using X-band (9.2 gHz) electron paramagnetic resonance (EPR) spectroscopy in aqueous solution at pH 7.4. In liquid solutions at 22°C, the vanadyl-protein complexes yielded VO2+ near rigid limit spectra. At 122 K, each of the three high-field resonances (i.e., 3/2, 5/2, and 7/2 parallel components) splits into two components indicating the presence of two classes of vanadyl-binding sites in each protein. The spectra of the frozen solutions were simulated to give parallel and perpendicular components of the hyperfine coupling constant and g factors similar to other vanadyl-protein complexes.  相似文献   

2.
Photosynthetically active reaction centre core (RCC) complexes were isolated from two species of green sulfur bacteria, Prosthecochloris (Ptc.) aestuarii strain 2K and Chlorobium (Chl.) tepidum, using the same isolation procedure. Both complexes contained the main reaction centre protein PscA and the iron–sulfur protein PscB, but were devoid of Fenna–Matthews–Olson (FMO) protein. The Chl. tepidum RCC preparation contained in addition PscC (cytochrome c). In order to allow accurate determination of the pigment content of the RCC complexes, the extinction coefficients of bacteriochlorophyll (BChl) a in several solvents were redetermined with high precision. They varied between 54.8 mM−1 cm−1 for methanol and 97.0 mM−1 cm−1 for diethylether in the QY maximum. Both preparations appeared to contain 16 BChls a of which two are probably the 132-epimers, 4 chlorophylls (Chls) a 670 and 2 carotenoids per RCC. The latter were of at least two different types. Quinones were virtually absent. The absorption spectra were similar for the two species, but not identical. Eight bands were present at 6 K in the BChl a QY region, with positions varying from 777 to 837 nm. The linear dichroism spectra showed that the orientation of the BChl a QY transitions is roughly parallel to the membrane plane; most nearly parallel were transitions at 800 and 806 nm. For both species, the circular dichroism spectra were dominated by a strong band at 807–809 nm, indicating strong interactions between at least some of the BChls. The absorption, CD and LD spectra of the four Chls a 670 were virtually identical for both RCC complexes, indicating that their binding sites are highly conserved and that they are an essential part of the RCC complexes, possibly as components of the electron transfer chain. Low temperature absorption spectroscopy indicated that typical FMO–RCC complexes of Ptc. aestuarii and Chl. tepidum contain two FMO trimers per reaction centre. This revised version was published online in June 2006 with corrections to the Cover Date.  相似文献   

3.
Absorption and CD spectra of bacteriochlorophyll a proteins and bacteriochlorophyll a reaction center complexes from two strains of Chlorobium limicola were recorded at 77 °K. Visual inspection showed that the Qy-band of chlorophyll in either protein was split into at least five components. Analysis of the spectra in terms of asymmetric Gaussian component pairs by means of computer program GAMET showed that six components are necessary to fit the spectra from strain 2K. These six components are ascribed to an exciton interaction between the seven bacteriochlorophyll a molecules in each subunit. The clear difference between the exciton splitting in the two bacteriochlorophyll a proteins shows that the arrangement of the chlorophyll molecules in each subunit must be slightly different.

The spectra for the bacteriochlorophyll a reaction center complexes have a component at 834 nm (absorption) and 832 nm (CD) which does not appear in the spectra of the bacteriochlorophyll a proteins. The new component is ascribed to a reaction center complex which is combined with bacteriochlorophyll a proteins to form the bacteriochlorophyll a reaction center complex. The complete absorption (or CD) spectrum for a given bacteriochlorophyll a reaction center complex can be described to a first approximation in terms of the absorption (or CD) spectrum for the corresponding bacteriochlorophyll a protein plus the new component ascribed to the reaction center complex.  相似文献   


4.
B C Sang  D M Gray 《Biochemistry》1989,28(24):9502-9507
Circular dichroism (CD) measurements were made on both fd and IKe gene 5 proteins in solution. The difference between the CD spectra of these two proteins was interpreted as being the result of an enhanced tyrosine contribution in the IKe gene 5 protein spectrum. There was no spectral evidence for significant alpha-helical structures in either of the two gene 5 proteins. CD measurements were also made on complexes of the two gene 5 proteins with poly(rA). The long-wavelength region (300-250 nm) of the CD spectra of both complexes was essentially like that of free poly(rA) at a high temperature. With the assumption that the poly(rA) components of the complexes had the same CD at all wavelengths as did free poly(rA) at a high temperature, it was possible to separate the CD spectra of the complexes into protein and nucleic acid components. Except for the tyrosine CD band at 229 nm, there were no significant changes in the CD bands of either protein upon binding to poly(rA). Thus, each protein appeared to maintain essentially the same overall secondary conformation when complexed with poly(rA) as when in its free state.  相似文献   

5.
X-band (9.2 GHz) electron spin resonance spectroscopy was used to investigate the binding of vanadyl to calmodulin. Solution spectra, obtained at ambient temperature with various VO2+:calmodulin molar ratios, suggested a binding stoichioimetry of 4 mol of VO2+/mol of protein and the possibility of two classes of binding sites. The latter was confirmed by using frozen solutions of calmodulin-VO2+ complexes that gave splitting of the spectral bands corresponding to the parallel components, which was particularly pronounced with the three high-field peaks. Competition of Ca2+ for the VO2+ binding sites was investigated, and the results indicated that two of the VO2+ sites corresponded to two of the Ca2+ sites; the other two VO2+ binding sites may have a higher affinity for VO2+ than for Ca2+ or they may correspond to Ca2+-independent sites. These results demonstrate that electron spin resonance spectroscopy can be used advantageously to probe subtle differences in the microenvironments of metal-binding sites in calmodulin.  相似文献   

6.
The formation of complex between the Cu2+ ion and poly(α-L -glutamic acid) [poly(Glu)] in 150 mM NaCl solutions was studied by uv–visible absorption and equilibrium dialysis methods at the mixing ratios of Glu residues to Cu2+, R, of 32, 16, and 8 and in the pH range 4–7. The results showed that more than 90% of Cu2+ ions bind to the poly(Glu) at pH > 4.9, but the bound Cu(II) begins to dissociate with a decrease in pH. The absorption spectra of bound Cu(II) varied with pH and R in a complicated manner. Three different component spectra were disclosed from the analysis of the pH dependence of the bound spectra. We concluded that poly(Glu)–Cu(II) complexes fall into three classes in the pH range 4–7, with the proportions of these complexes varying with both pH and R. The three complexes predominate either in the helix or extended-coil region, in the helix–coil transition region, or in the helix-aggregate region. The stability constant and binding mode of each Cu(II)–Glu complex were estimated from the dialysis data. With these results, the possible structure of each complex is discussed.  相似文献   

7.
The structure and speciation of the complexes formed between mercury(II) ions and glutathione (GSH = L-glutamyl-L-cysteinyl-glycine) have been studied for a series of alkaline aqueous solutions (\( C_{{{\text{Hg}}^{{2 + }}}}\,{\sim18\,{\rm{mmol}}\,{\rm{{dm^{-3}}}}}\) and C GSH = 40–200 mmol dm?3 at pH ~10.5) by means of extended X-ray absorption fine structure (EXAFS) and 199Hg NMR spectroscopy at ambient temperature. The dominant complexes are [Hg(GS)2]4? and [Hg(GS)3]7?, with mean Hg–S bond distances of 2.32(1) and 2.42(2) Å observed in digonal and trigonal Hg–S coordination, respectively. The proportions of the Hg2+–glutathione complexes were evaluated by fitting linear combinations of model EXAFS oscillations representing each species to the experimental EXAFS spectra. The [Hg(GS)4]10? complex, with four sulfur atoms coordinated at a mean Hg–S bond distance of 2.52(2) Å, is present in minor amounts (<30%) in solutions containing a large excess of glutathione (C GSH ≥ 160 mmol dm?3). Comparable alkaline mercury(II) cysteine (H2Cys) solutions were also investigated and a reduced tendency to form higher complexes was observed, because the deprotonated amino group of Cys2? allows the stable [Hg(S,N-Cys)2]2? chelate to form. The effect of temperature on the distribution of the Hg2+–glutathione complexes was studied by comparing the EXAFS spectra at ambient temperature and at 25 K of a series of glycerol/water (33/67, v/v) frozen glasses with \( C_{{{\text{Hg}}^{{2 + }} }} \,{\sim7\,{\rm{mmol}}\,{\rm{{dm^{-3}}}}} \) and C GSH = 16–81 mmol dm?3. Complexes with high Hg–S coordination numbers, [Hg(GS)3]7? and [Hg(GS)4]10?, became strongly favored when just a moderate excess of glutathione (C GSH ≥28 mmol dm?3) was used in the glassy samples, as expected for a stepwise exothermic bond formation. Addition of glycerol had no effect on the Hg(II)–glutathione speciation, as shown by the similarity of the EXAFS spectra obtained at room temperature for two parallel series of Hg(II)-glutathione solutions with \( C_{{{\text{Hg}}^{{2 + }} }} \,{\sim7\,{\rm{mmol}}\,{\rm{{dm^{-3}}}}},\) with and without 33% glycerol. Also, the 199Hg NMR chemical shifts of a series of ~18 mmol dm?3 mercury(II) glutathione solutions with 33% glycerol were not significantly different from those of the corresponding series in aqueous solution.  相似文献   

8.
A. F. Bunkin 《Biophysics》2012,57(6):709-715
The novel method of nonlinear laser spectroscopy — low frequency spectroscopy of four-photon scattering of laser radiation was applied to detect a considerable growth of ortho-H2O spin isomer and also H2O2 molecule concentration in a hydrate layer at the interface between water and DNA, denatured DNA molecules and α-chymotrypsin. Spectra of rotational resonances of ortho/para-H2O spin isomers were observed in aqueous solutions of different biopolymers and also in distilled water in the range from zero to 100 cm?1 with the spectral resolution of 0.05–0.1 cm?1. The fitting of four-wave mixing spectra shows notable growth of the H2O2 concentration and rotational line’s amplitude by a factor of ~3 in DNA solutions due to denaturizing. Besides, we studied the four-photon scattering spectra of α-chymotrypsin aqueous solutions at protein concentrations between 0 and 20 mg/cm3 in the range of ±7 cm?1. We found that the velocity of sound in the protein aqueous solution measured by the shift of the Mandelstam-Brillouin scattering spectrum components was a cubic dependence on the protein concentration and reached the value of about 3000 m/s at 20 mg/cm3.  相似文献   

9.
We have studied the electron paramagnetic resonance (epr) spectra of complexes of apo-yeast enolase with 65Cu+2 in the presence and absence of substrate and magnesium ion. An unusual epr spectrum with large g parallel, large g and A rhombicity and very narrow line-widths (10 G) is seen for the first two 65Cu+2 bound in the presence of substrate 2-phosphoglycerate (2PGA). the epr parameters, consistent with rhombic and tetragonal distortion of an octahedral geometry of the coordination sphere of the Cu+2 are g = (2.123, 2.042, 2.405) and A = (2.58, 4.19, 12.0) mK. The high g parallel and absence of super-hyperfine splitting are strong evidence for absence of nitrogen ligands. In the presence of Mg+2 and 2PGA, the Cu+2-enolase solutions exhibit a complex epr spectrum reflecting exchange and dipolar interaction between the first two Cu+2 ions bound. The spectra of Cu+2 plus enolase in the presence and absence of Mg+2 without 2PGA are distinct but not unambiguous, each reflecting at least two inequivalent binding sites. In addition to providing information on the geometry and location of the divalent cation binding sites, the data show unequivocally that imidazole residues, previously found to have a role in catalysis, do not participate in Cu+2 binding. Although Cu+2 does not activate the enzyme, direct binding measurements show that Cu+2 competes stoichiometrically with the activating ion, Mg+2. A reinterpretation of earlier Mn+2 enolase studies is proposed to reconcile the Cu+2 and Mn+2 data.  相似文献   

10.
《Inorganica chimica acta》2007,360(9):2973-2982
Syntheses and crystal structures of two molecular, heteroleptic cadmium complexes with CdS2NO2 and CdS2N2 kernels are described. Bis(tri-tert-butoxysilanethiolate)(1-methylimidazole)cadmium(II) and bis(tri-tert-butoxysilanethiolate)bis(1-methylimidazole)cadmium(II) coexist at equilibrium in chloroform solutions with varying concentrations of bis[bis(tri-tert-butoxysilanethiolate)cadmium(II)] and 1-methylimidazole. The equilibrium is characterized by solution 113Cd NMR spectra. Solid state CP MAS 13C, 29Si, 113Cd NMR data for the complexes are also reported, analyzed and compared with the results obtained for cadmium-substituted proteins. The similarities and differences between the structures of cadmium complexes and their zinc analogues are discussed.  相似文献   

11.
The ferriheme resonances of the low-spin (S = 1/2) complexes of wild-type (wt) nitrophorin 2 (NP2) and its heme pocket mutant NP2(V24E) with imidazole (ImH), histamine (Hm), and cyanide (CN) as the sixth ligand have been investigated by NMR spectroscopy as a function of pH (4.0–7.5). For the three wt NP2 complexes, the ratio of the two possible heme orientational isomers, A and B, remains almost unchanged (ratio of A:B approximately 1:6 to 1:5) over this wide pH range. However, strong chemical exchange cross peaks appear in the nuclear Overhauser effect spectroscopy/exchange spectroscopy (NOESY/EXSY) spectra for the heme methyl resonances at low pH (pH* 4.0–5.5), which indicate chemical exchange between two species. We have shown these to be two different exogenous ImH or Hm orientations that are denoted B and B′, with the ImH plane nearly parallel and perpendicular to the ImH plane of the protein-provided His57, respectively. The wt NP2–CN complex also shows EXSY cross peaks due to chemical exchange, which is shown to be a result of interchange between two ruffling distortions of the heme. The same ruffling distortion interchange is also responsible for the ImH and Hm chemical exchange. For the three NP2(V24E) ligand complexes, no EXSY cross peaks are observed, but the A:B ratios change dramatically with pH. The fact that heme favors the A orientation highly for NP2(V24E) at low pH as compared with wt NP2 is believed to be due to the steric effect of the V24E mutation. The existence of the B′ species at lower pH for wt NP2 complexes and the increase in A heme orientation at lower pH for NP2(V24E) are believed to be a result of a change in structure near Glu53 when it is protonated at low pH. 1H{13C} heteronuclear multiple quantum coherence (HMQC) spectra are very helpful for the assignment of heme and nearby protein side chain resonances.  相似文献   

12.
In aqueous solutions of NAD(H), there is an equilibrium between two different conformations : a “folded” conformation in which adenine and nicotinamide are staked together and an “unfolded” conformation in which the two rings are without interaction.The folded conformation is the more stable in aqueous solution whereas in organic solution it is the unfolded one.As we have previously shown, the PMR spectra of Co2+—NAD(H) complexes may be related with the coenzyme conformation giving suggest to a new method for NAD(H) conformational analysis.The results of this method applied to methanol 2H2O and dioxane/2H2O solutions are reported in this paper: they are in good accordance with those of spectrofluorimetric analysis.  相似文献   

13.
The B800–850 antenna complex of Rhodopseudomonas sphaeroides was studied by comparing the spectral properties of several different types of complexes, isolated from chromatophores by means of the detergents lithium dodecyl sulfate (LDS) or lauryl dimethylamine N-oxide (LDAO). Fluorescence polarization spectra of the BChl 800 emission at 4 K indicated that rapid energy transfer between at least two BChl 800 molecules occurs with a rate constant of energy transfer kET > 3 · 1012 s?1. The maximal dipole-dipole distance between the two BChl 800 molecules was calculated to be 18–19 Å. The porphyrin rings of the BChl 800 molecules are oriented parallel to each other, while their Qy transition moments are mutually perpendicular. The energy-transfer efficiency from carotenoid to bacteriochlorophyll measured in different complexes showed that two functionally different carotenoids are present associated with, respectively, BChl 800 and BChl 850. Fluorescence polarization and linear dichroism spectra revealed that these carotenoids have different absorption spectra and a different orientation with respect to the membrane. The carotenoid associated with BChl 800 absorbs some nanometers more to the red and its orientation is approximately parallel to the membrane, while the carotenoid associated with BChl 850 is oriented more or less perpendicular to the membrane. The fluorescence polarization of BChl 850 was the same for the different complexes. This indicates that the observed polarization of the fluorescence is determined by the smallest complex obtained which contains 8–10 BChl 850 molecules. The B800–850 complex isolated with LDAO thus must consist of a highly ordered array of smaller structures. On basis of these results a minimal model is proposed for the basic unit consisting of four BChl 850 and two BChl 800 and three carotenoid molecules.  相似文献   

14.
We have studied the molybdenum-iron protein (MoFe protein, also known as component I) from Azobacter vinelandi using M?ssbauer spectroscopy and electron paramagnetic resonance on samples enriched with 57Fe. These spectra can be interpreted in terms of two EPR active centers, each of which is reducible by one electron. A total of four different chemical environments of Fe can be discerned. One of them is a cluster of Fe atoms with a net electronic spin of 3/2, one of them is high-spin ferrous iron and the remaining two are iron in a reduced state (probably in clusters). The results are as follows: Chemical analysis yields 11.5 Fe atoms and 12.5 labile sulfur atoms per molybdenum atom; the molecule contains two Mo atoms per 300 000 daltons. The EPR spectrum of the MoFe protein exhibits g values at 4.32, 3.65 and 2.01, associated with the ground state doublet of a S = 3/2 spin system. The spin Hamiltonian H = D(S2/z minus 5/4 + lambda(S2/x minus S2/y)) + gbeta/o S-H fits the experimental data for go = 2.00 and lambda = 0.055. Quantitative analysis of the temperature dependence of the EPR spectrum yields D/k = 7.5 degrees K and 0.91 spins/molybdenum atom, which suggests that the MoFe protein has two EPR active centers. Quantitative evaluation of M?ssbauer spectra shows that approximately 8 iron atoms give rise to one quadrupole doublet; at lower temperatures magnetic spectra, associated with the groud electronic doublet, are observed; at least two magnetically inequivalent sites can be distinguished. Taken together the data suggest that each EPR center contains 4 iron atoms. The EPR and M?ssbauer data can only be reconciled if these iron atoms reside in a spin-coupled (S = 3/2) cluster. Under nitrogen fixing conditions the magnetic M?ssbauer spectra disappeared concurrently with the EPR signal and quadrupole doublets are obserced at all temperatures. The data suggest that each EPR active center is reduced by one electron. The M?ssbauer investigation reveals three other spectral components characteristic of iron nuclei in an environment of integer or zero electronic spin, i.e. they reside in complexes which are "EPR-silent". One of the components (3-4 iron atoms) has M?ssbauer parameters characteristic of the high-spin ferrous iron as in reduced ruberdoxin. However, measurements in strong fields indicate a diamagnetic environment. Another component, representing 9-11 iron atoms, seems to be diamagnetic also. It is suggested that these atoms are incorporated in spin-coupled clusters.  相似文献   

15.
The phototransformation of protochlorophyllide forms was studied in epicotyls of dark-germinated pea (Pisum sativum L. cv. Zsuzsi) seedlings. Middle segments were illuminated with white or 632.8 nm laser flash or continuous light at room temperature and at −15°C. At low light intensities, photoreduction could be distinguished from bleaching. 77 K fluorescence emission spectra were measured, difference spectra of illuminated and non-illuminated samples were calculated and/or the spectra were deconvoluted into Gaussian components. The 629 nm-emitting protochlorophyllide form, P629 (Pxxx where xxx is the fluorescence emission maximum), was inactive. For short-period (2–100 ms) and/or low-intensity (0.75–1.5 µmol m−2 s−1) illumination, particularly with laser light, the transformation of P636 into the 678 nm-emitting chlorophyllide form, C678 (Cxxx where xxx is the fluorescence emission maximum), was characteristic. This process was also found when the samples were cooled to −15°C. The transformation of P644 into C684 usually proceeded in parallel with the process above as a result of the strong overlap of the excitation bands of P636 and P644. The Shibata shift of C684 into a short-wavelength form, C675–676, was observed. Long-period (20–600 s) and/or high-intensity (above 10 µmol m−2 s−1) illumination resulted in the parallel transformation of P655 into C692. These results demonstrate that three flash-photoactive protochlorophyllide forms function in pea epicotyls. As a part of P636 is flash photoactive, its protochlorophyllide molecule must be bound to the active site of a monomer protein unit [Böddi B, Kis-Petik K, Kaposi AD, Fidy J, Sundqvist C (1998) The two short wavelength protochlorophyllide forms in pea epicotyls are both monomeric. Biochim Biophys Acta 1365: 531–540] of the NADPH:protochlorophyllide oxidoreductase (EC 1.3.1.33). Dynamic interconversions of the protochlorophyllide forms into each other, and their regeneration, were also found, which are summarized in a scheme.  相似文献   

16.
The reactions of [Ru(PPh3)3Cl2], N-(benzoyl)-N′-(5-R-salicylidene)hydrazines (H2bhsR, R = H, OCH3, Cl, Br and NO2) and triethylamine (1:1:2 mole ratio) in methanol afford mononuclear ruthenium(III) complexes having the general formula trans-[Ru(bhsR)(PPh3)2Cl]. In the case of R = H, a dinuclear ruthenium(III) complex of formula [Ru2(μ-OCH3)2(bhsH)2(PPh3)2] has been isolated as a minor product. The complexes are characterized by elemental analysis, magnetic, spectroscopic and electrochemical measurements. The crystal structures of the dinuclear complex and two mononuclear complexes have been determined. In the dinuclear complex, each metal centre is in distorted octahedral NO4P coordination sphere constituted by the two bridging methoxide groups, one PPh3 molecule and the meridionally spanning phenolate-O, imine-N and amide-O donor bhsH2−. The terminal PPh3 ligands are trans to each other. In the mononuclear complexes, bhsR2− and the chlorine atom form an NO2Cl square-plane around the metal centre and the P-atoms of the two PPh3 molecules occupy the remaining two axial sites to complete a distorted octahedral NO2ClP2 coordination sphere. All the complexes display ligand-to-metal charge transfer bands in the visible region of the electronic spectra. The cryomagnetic measurements reveal the antiferromagnetic character of the diruthenium(III) complex. The low-spin mononuclear ruthenium(III) complexes as well as the diruthenium(III) complex display rhombic EPR spectra in frozen solutions. All the complexes are redox active in CH2Cl2 solutions. Two successive metal centred oxidations at 0.69 and 1.20 V (versus Ag/AgCl) are observed for the dinuclear complex. The mononuclear complexes display a metal centred reduction in the potential range −0.53 to −0.27 V. The trend in these potential values reflects the polar effect of the substituents on the salicylidene moiety of the tridentate ligand.  相似文献   

17.
ESR spectra of the tight binding Cu(II) complex of bovine serum albumin (BSA) has been studied using S-band. At physiological pH, only one form of copper binding to BSA was detected from the ESR spectra. From previous X-band ESR spectra, nitrogen superhyperfine splittings were observable in the g perpendicular region; however, the resolution of the g parallel region was not sufficient to confirm the exact donor atoms of the complex. Using low-frequency ESR (2-4 GHz) at 77 K, we have resolved the nitrogen superhyperfine structure in the g parallel region. A computer simulation method has been developed for distinguishing between three and four nitrogen donor atoms. The Hyde-Froncisz theory of g and A strain broadening has been modified to use a field-swept calculation for the line shape. The observed intensity pattern and the computer simulation of such spectra positively confirm the structure of Cu(II) ion coordinated to four in-plane nitrogen atoms in frozen aqueous solutions of Cu(II)-BSA complexes at physiological pH. This is the first time that this binding site has been confirmed on the protein instead of a protein fragment or model compound. This work is another example of the usefulness of the S-band ESR technique for characterizing the metal-protein interactions when random variation in g factors cause line broadening in conventional X-band ESR spectra.  相似文献   

18.
Core antenna and reaction centre of photosytem I (PS I) complexes from the cyanobacteria Arthrospira platensis and Thermosynechococcus elongatus have been characterized by steady-state polarized absorption spectroscopy, including linear dichroism (LD) and circular dichroism (CD). CD spectra and the second derivatives of measured 77 K CD spectra reveal the spectral components found in the polarized absorption spectra indicating the excitonic origin of the spectral forms of chlorophyll in the PS I complexes. The CD bands at 669-670(+), 673(+), 680(−), 683-685(−), 696-697(−), and 711(−) nm are a common feature of used PSI complexes. The 77 K CD spectra of the trimeric PS I complexes exhibit also low amplitude components around 736 nm for A. platensis and 720 nm for T. elongatus attributed to red-most chlorophylls. The LD measurements indicate that the transition dipole moments of the red-most states are oriented parallel to the membrane plane. The formation of P700+A1 or 3P700 was monitored by time-resolved difference absorbance and LD spectroscopy to elucidate the spectral properties of the PS I reaction centre. The difference spectra give strong evidence for the delocalization of the excited singlet states in the reaction centre. Therefore, P700 cannot be considered as a dimer but should be regarded as a multimer of the six nearly equally coupled reaction centre chlorophylls in accordance with structure-based calculations. On the basis of the results presented in this work and earlier work in the literature it is concluded that the triplet state is localized most likely on PA, whereas the cation is localized most likely on PB.  相似文献   

19.
The synthesis and characterization of the block copolypeptide (Leu67, Lys33)100Orn20, a synthetic model of histone, are reported. In neutral aqueous solutions, 80% of the etheropolypeptide block assumes an α-helical conformation, whereas the polyornithine block is in a random-coil conformation. In the association complexes with DNA, melting and titration experiments, as well as CD results, indicate that the polyornithine block interacts with DNA, whereas at least 2/3 of the lysine residues of the (Leu, Lys) moiety are excluded from the direct binding with DNA. CD spectra of the association complexes reveal significant differences from those obtained with DNA–polyornithine and DNA–polylysine complexes but substantial similarities with CD spectra of native and reconstituted nucleohistones. In contrast to DNA–polyornithine complexes, the CD spectra of the ternary complexes, copolypeptide–DNA–ethidium bromide, indicate a strong reduction of the dye intercalation. The low-angle x-ray diffraction pattern, reminiscent of that of chromatin, reveals the presence of a superstructure in these complexes. The results obtained are discussed in connection with the expected structural features of the model.  相似文献   

20.
Examination of the secondary structure of proteins by deconvolved FTIR spectra   总被引:70,自引:0,他引:70  
D M Byler  H Susi 《Biopolymers》1986,25(3):469-487
Fourier transform ir (FTIR) spectra of 21 globular proteins have been obtained at 2 cm?1 resolution from 1600 to 1700 cm?1 in deuterium oxide solution. Fourier self-deconvolution was applied to all spectra, revealing that the amide I band of each protein except casein consists of six to nine components. The components are observed at 11 well-defined frequencies, although all proteins do not exhibit components at every characteristic frequency. The root mean square (RMS) deviation of 124 individual values from the 11 average characteristic frequencies is 1.9 cm?1. The observed components are assigned to helical segments, extended beta-segments, unordered segments, and turns. Segments with similar structures do not necessarily exhibit band components with identical frequencies. For instance, the lower frequency beta-structure band can vary within a range of approximately 15 cm?1. The relative areas of the individual components of the deconvolved spectra were determined by a Gauss–Newton, iterative curve-fitting procedure that assumed Gaussian band envelopes for the deconvolved components. The measured areas were used to estimate the percentage of helix and beta-structure for each of 21 globular proteins. The results are in good general agreement with values derived from x-ray data by Levitt and Greer. The RMS deviation between 22 values (alpha- and beta-content of 11 beta-rich proteins measured by both techniques) is 2.5 percentage points; the maximum absolute deviation is 4 percentage points.  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