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 interaction between the 15-mer oligonucleotide d[A(TA)(7)].d[T(AT)(7)] and the three biogenic polyamines, putrescine, spermidine and spermine, under physiological conditions has been studied by Raman spectroscopy. Solutions containing 60 mM (phosphate) of the oligonucleotide and different polyamine concentrations ranging from 1 to 75 mM have been studied. Both natural and heavy water were used as solvents. Difference Raman spectra were computed by subtracting the sum of the separated component spectra from the experimental spectra of the complexes. The Raman data suggested that the interaction of biogenic polyamines with d[A(TA)(7)].d[T(AT)(7)] presents differences related with their sizes and electric charges. Preferential bindings through the oligonucleotide minor groove for putrescine and spermidine were proposed. Spermine would interact by both minor and major grooves, although interaction by the minor groove seems to be more favored. Main reactive sites were thymine-O2 and adenine-N3 atoms at the minor grooves and adenine-N7 and thymine-O4 at the major groove. Electrostatic attractions between the polyamine amino and oligonucleotide phosphodioxy groups were also proposed. Under our experimental conditions, no macromolecular effects on d[A(TA)(7)].d[T(AT)(7)] (changes on secondary or tertiary structures) were detected from Raman spectroscopy, contrary to what happened for GC sequences at the same experimental settings. This fact agrees with the role of the biogenic polyamines during the first steps of the macromolecular synthesis, which involve DNA opening in AT motifs.  相似文献   

2.
The interaction between the oligonucleotide d[G(CG)(7)]. d[C(GC)(7)] and the three biogenic polyamines putrescine, spermidine, and spermine under physiological conditions has been studied by Raman spectroscopy. The results indicate the formation of highly ordered aggregated structures in solution, largely stabilized by electrostatic attractions, which have been described as cholesteric phases. Aggregation seems to be preceded by a partial B --> Z conformational transition for spermidine and spermine, which would allow for a deeper oligonucleotide-polyamine interaction. Interaction with the nucleic bases has also been evidenced for aggregates. At low polyamine concentrations the preferential binding sites are similar to those proposed for their interactions with ct-DNA. With increasing the polyamine concentration, the oligonucleotide-polyamine interactions involve both minor and major grooves, which is consistent with the formation of cholesteric phases.  相似文献   

3.
Biogenic polyamines, such as putrescine, spermidine, and spermine are small organic polycations involved in numerous diverse biological processes. These compounds play an important role in nucleic acid function due to their binding to DNA and RNA. It has been shown that biogenic polyamines cause DNA condensation and aggregation similar to that of inorganic cobalt(III)hexamine cation, which has the ability to induce DNA conformational changes. However, the nature of the polyamine.DNA binding at the molecular level is not clearly established and is the subject of much controversy. In the present study the effects of spermine, spermidine, putrescine, and cobalt(III)hexamine on the solution structure of calf-thymus DNA were investigated using affinity capillary electrophoresis, Fourier transform infrared, and circular dichroism spectroscopic methods. At low polycation concentrations, putrescine binds preferentially through the minor and major grooves of double strand DNA, whereas spermine, spermidine, and cobalt(III)hexamine bind to the major groove. At high polycation concentrations, putrescine interaction with the bases is weak, whereas strong base binding occurred for spermidine in the major and minor grooves of DNA duplex. However, major groove binding is preferred by spermine and cobalt(III)hexamine cations. Electrostatic attractions between polycation and the backbone phosphate group were also observed. No major alterations of B-DNA were observed for biogenic polyamines, whereas cobalt(III)hexamine induced a partial B --> A transition. DNA condensation was also observed for cobalt(III)hexamine cation, whereas organic polyamines induced duplex stabilization. The binding constants calculated for biogenic polyamines are K(Spm) = 2.3 x 10(5) M(-1), K(Spd) = 1.4 x 10(5) M(-1), and K(Put) = 1.02 x 10(5) M(-1). Two binding constants have been found for cobalt(III)hexamine with K(1) = 1.8 x 10(5) M(-1) and K(2) = 9.2 x 10(4) M(-1). The Hill coefficients indicate a positive cooperativity binding for biogenic polyamines and a negative cooperativity for cobalt(III)hexamine.  相似文献   

4.
Fourier transform infrared and laser Raman spectroscopies were used to study the effects of dodecylpyridinium bromide on the conformation of haemoglobin, myoglobin, bovine serum albumin, ribonuclease, ovalbumin, lysozyme, trypsin and beta-lactoglobulin in aqueous solution. Addition of the cationic detergent caused a decrease in alpha-helix conformation in highly helical proteins. At low detergent concentrations stabilization of beta-sheet conformation was observed.  相似文献   

5.
Biogenic polyamines, putrescine, spermidine, and spermine, are ubiquitous cellular cations and exert multiple biological functions. Polyamine analogues mimic biogenic polyamines at macromolecular level but are unable to substitute for natural polyamines and maintain cell proliferation, indicating biomedical applications. The mechanistic differences in DNA binding mode between natural and synthetic polyamines have not been explored. The aim of this study was to examine the interaction of calf thymus DNA with three polyamine analogues, 1,11-diamino-4,8-diazaundecane (333), 3,7,11,15-tetrazaheptadecane x 4 HCl (BE-333), and 3,7,11,15,19-pentazahenicosane x 5 HCl (BE-3333), using FTIR, UV-visible, and CD spectroscopy. Polyamine analogues bind with guanine and backbone PO2 group as major targets in DNA, whereas biogenic polyamines bind to major and minor grooves as well as to phosphate groups. Weaker interaction with DNA was observed for analogues with respect to biogenic polyamines, with K(333) = 1.90 (+/-0.5) x 10(4) M(-1), K(BE-333) = 6.4 (+/-1.7) x 10(4) M(-1), K(BE-3333) = 4.7 (+/-1.4) x 10(4) M(-1) compared to K(Spm) = 2.3 (+/-1.1) x 10(5) M(-1), K(Spd) = 1.4 (+/-0.6) x 10(5) M(-1), and K(Put) = 1.02 (+/-0.5) x 10(5) M(-1). A partial B- to A-DNA transition was also provoked by analogues. These data suggest distinct differences in the binding of natural and synthetic polyamines with DNA.  相似文献   

6.
Shi YB  Fang JL  Liu XY  Du L  Tang WX 《Biopolymers》2002,65(2):81-88
The secondary structures of porcine brain Cu(4)Zn(3)-metallothionein (MT)-III and Cd(5)Zn(2)MT-I, Cd(5)Zn(2)MT-II, and Zn(7)MT-I from rabbit livers in the solid state are investigated by Fourier transform IR spectroscopy (FTIR) and Fourier transform Raman spectroscopy (FT-Raman). The Cu(4)Zn(3)MT-III contains 26-28% beta-turns and half-turns, 13-14% 3(10)-helices, 47-49% random coils, and 11-12% beta-extended chains. The structural comparison of porcine brain Cu(4)Zn(3)MT-III with rabbit liver Cd(5)Zn(2)MT-I (II) and Zn(7)MT-I shows that the contents of the random coil structure are obviously increased. The results indicate that the insert of an acidic hexapeptide in the alpha domain of Cu(4)Zn(3)MT-III possibly forms an alpha helix. However, because the bands assigned to the alpha-helix and random coil structures are overlapped in the spectra, the content of random coil structures in Cu(4)Zn(3)MT-III is therefore higher than those in Cd(5)Zn(2)MT-I, Cd(5)Zn(2)MT-II, and Zn(7)MT-I.  相似文献   

7.
8.
The Pr and Pfr forms of phytochrome in H2O and D2O have been studied by Fourier transform resonance Raman spectroscopy with near-infrared excitation (1064 nm). It is demonstrated that this technique is a powerful method for analyzing the chromophore structures of photosensitive pigments. The high spectral quality allows discussion of vibrational assignments based on an empirical approach using previously published data obtained from model compounds. The reduction in intensity of a high-frequency band assigned to the ring-C/D methine bridge vibration is an indication for the non-coplanarity of the ring D in Pfr. The high intensity of a C-H out-of-plane vibration also supports this hypothesis. In Pr, a broad peak at approximately 1100 cm-1 is assigned to an out-of-plane vibration of a strongly hydrogen-bonded pyrrole C=NH+ group. It is missing in Pfr, suggesting deprotonation of the corresponding ring during the transformation from Pr to Pfr.  相似文献   

9.
Fourier transform infrared spectroscopy was used to investigate the small conformational differences which exist between ribonuclease A and ribonuclease S in aqueous systems. Deconvolution and derivative methods were used to observe the overlapping components of the amide I and II bands. These proteins give identical spectra in H2O and after complete exchange in 2H2O. However structural differences are revealed by monitoring the rate of 1H-2H exchange by Fourier transform infrared spectroscopy. At equivalent times of exposure in 2H2O buffer ribonuclease S undergoes greater isotopic exchange than ribonuclease A. Thus complete exchange takes place for ribonuclease S but not ribonuclease A after incubation at room temperature for 8 days. Complete 1H-2H exchange of ribonuclease A was achieved by incubation at 62 degrees C for 30 min. The available X-ray data and comparison with the infrared spectra of other soluble proteins was used to assign the components of the amide I and II bands to various secondary structures. In particular, band shifts observed during the later stages of exchange are associated with slowly exchanging residues in beta-strand and alpha-helical regions. The higher rate of exchange for ribonuclease S is associated with a greater conformational flexibility and a more open structure. The results show that it is necessary to be cautious in making band assignments based on exchange methods unless the extent of exchange is known. Furthermore, it is seen that the combination of Fourier transform infrared spectroscopy and hydrogen-deuterium exchange is a powerful technique for revealing small differences in protein secondary structure.  相似文献   

10.
Hemolysis is a very common phenomenon and is referred as the release of intracellular components from red blood cells to the extracellular fluid. Hemolyzed samples are often rejected in clinics due to the interference of hemoglobin and intracellular components in laboratory measurements. Plasma and serum based vibrational spectroscopy studies are extensively applied to generate spectral biomarkers for various diseases. However, no studies have reported the effect of hemolysis in blood based vibrational spectroscopy studies. This study was undertaken to evaluate the effect of hemolysis on infrared and Raman spectra of blood plasma. In this study, prostate cancer plasma samples (n = 30) were divided into three groups (nonhemolyzed, mildly hemolyzed, and moderately hemolyzed) based on the degree of hemolysis and FTIR and Raman spectra were recorded using high throughput (HT)‐FTIR and HT‐Raman spectroscopy. Discrimination was observed between the infrared and Raman spectra of nonhemolyzed and hemolyzed plasma samples using principal component analysis. A classical least square fitting analysis showed differences in the weighting of pure components in nonhemolyzed and hemolyzed plasma samples. Therefore, it is worth to consider the changes in spectral features due to hemolysis when comparing the results within and between experiments.  相似文献   

11.
The effect of a series of aminooxy analogues of the biogenic polyamines spermidine and spermine on the conformation of calf thymus DNA is studied. These new molecules are isosteric and charge insufficient analogues that are suitable to study the roles of both charge distribution and structural requirements in the molecular physiology of the biogenic polyamines. They are also evidenced as useful tools to inhibit polyamine biosynthesis and cell growth. Circular dichroism (CD) spectra of solutions containing DNA and the aminooxy analogues at different concentrations (100-1000 microM) and different pH values, (5-7.5) are recorded. We use both sonicated and highly polymerized calf thymus DNA. The CD spectra of sonicated DNA showed the formation of Psi-DNA, a highly ordered aggregated structure similar to liquid crystals, in the presence of the aminooxy analogues. Aggregation induced by an aminooxy derivative of spermine is followed by DNA collapse when increasing the polyamine concentration. The features of Psi-DNA are not detected for highly polymerized DNA. Temperature melting measurements support a high degree of structural order of the aggregates. The CD experiments indicate that dications are unable to induce major changes on the macromolecular structure of DNA. In addition, aggregation is only observed when the trimethylene moiety is present between two adjacent positive charges. The observed differences among the CD spectra of DNA solutions with different aminooxy derivatives of spermidine indicate different roles for different amino groups of this biogenic polyamine when interacting with DNA. Our results support the idea that aminooxy analogues can be used as good models in studying the physiological functions of biogenic polyamines.  相似文献   

12.
The interaction between a variety of polyamines, both naturally occurring and synthetic, and calf thymus DNA has been studied using 23Na NMR. The relaxation behaviour of 23Na reflects the extent of interaction of Na+ with DNA phosphate groups and therefore the extent of charge neutralisation of DNA phosphate groups (P) by polyamine amino and imino groups (N) in solutions of DNa, polyamine and Na+. The studies reveal that whereas spermine and spermidine are capable of expelling nearly all of the Na+ ions from DNA at N/P approximately 1, diamines such as putrescine and homologues of spermine and spermidine are capable of neutralising only roughly 50% of DNA phosphates. The results provide a challenge to current models of DNA-polyamine interactions.  相似文献   

13.
W L Chen  S M Nie  J F Kuck  Jr    N T Yu 《Biophysical journal》1991,60(2):447-455
We present in this report a detailed structural study of calf gamma-crystallins both in the solid state and in solution by the newly developed technique of near-infrared (IR) Fourier transform (FT)-Raman spectroscopy as well as by the conventional Raman method. In comparison with conventional laser Raman spectroscopy, the near-IR FT-Raman approach exhibits several attractive features such as fluorescence rejection capability, frequency accuracy, and the FT's multiplex and throughput advantages. These distinct characteristics combined form the basis for the particular suitability of FT-Raman in crystallin structural analysis and elucidation. We have thus obtained evidence in support of the view that native calf gamma-II crystallin does not contain a disulfide bond either in the lyophilized state or in solution. In addition, conventional Raman spectra are examined for all four gamma-crystallin fractions. gamma-S, gamma-II, gamma-III, and gamma-IV, and the results indicate a high degree of structural similarities among them. It is also found that the sulfhydryl groups in all four gamma-crystallins are highly resistant to air oxidation and are capable of maintaining their reduced state during isolation in the absence of added reductants or such chelating agents as EDTA.  相似文献   

14.
15.
C Otto  F F de Mul  J Greve 《Biopolymers》1987,26(10):1667-1689
Raman spectra of the bacteriophage T4 denaturing protein gp32, its complex with the polynucleotides poly(rA), poly(dA), poly(dT), poly(rU), and poly(rC), and with the oligonucleotides (dA)8 and (dA)2, were recorded and interpreted. According to an analysis of the gp32 spectra with the reference intensity profiles of Alix and co-workers [M. Berjot, L. Marx, and A. J. P. Alix (1985) J. Ramanspectrosc., submitted; A. J. P. Alix, M. Berjot, and J. Marx (1985) in Spectroscopy of Biological Molecules, A. J. P. Alix, L. Bernard, and M. Manfait, Eds., pp. 149–154], 1 gp32 contains ≈ 45% helix, ≈ 40% β-sheet, and 15% undefined structure. Aggregation of gp32 at concentrations higher than 40 mg/mL leads to a coordination of the phenolic OH groups of 4–6 tyrosines and of all the sulfhydryl (SH) groups present in the protein with the COO? groups of protein. The latter coordination persists even at concentrations as low as 1 mg/mL. In polynucleotide–protein complexes the nucleotide shields the 4–6 tyrosine residues from coordination by the COO? groups even at high protein concentration. The presence of the nucleotide causes no shielding of the SH groups. With Raman difference spectroscopy it is shown that binding of the protein to a single-stranded nucleotide involves both tyrosine and trytophan residues. A change in the secondary structure of the protein upon binding is observed. In the complex, gp32 contains more β-sheet structure than when uncomplexed. A comparison of the spectra of complexed poly(rA) and poly(dA) with the spectra of their solution conformations at 15°C reveals that in both polynucleotides the phosphodiester vibration changes upon complex formation in the same way as upon a transition from a regular to a more disordered conformation. Distortion of the phosphate–sugar–base conformation occurs upon complex formation, so that the spectra of poly(rA) and poly(dA) are more alike in the complex than they are in the free polynucleotides. The decrease in intensity of the Raman bands at 1304 cm?1 in poly(rA), at 1230 cm?1 in poly(rU), and at 1240 and 1378 cm?1 of poly(dT) may be indicative of increased stacking interactions in the complex. No influence of the nucleotide chain length upon the Raman spectrum of gp322 in the complex was detected. Both the nucleotide lines and the protein lines in the spectrum of a complex are identical in poly(dA) and (dA)8.  相似文献   

16.
Polyamines such as spermidine and spermine are abundant in living cells and are believed to aid in the dense packaging of cellular DNA. DNA condensation is a prerequisite for the transport of gene vectors in living cells. To elucidate the structural features of polyamines governing DNA condensation, we studied the collapse of lambda-DNA by spermine and a series of its homologues, H2N(CH2)3NH(CH2)n=2-12NH(CH2)3NH2 (n = 4 for spermine), using static and dynamic light scattering techniques. All polyamines provoked DNA condensation; however, their efficacy varied with the structural geometry of the polyamine. In 10 mM sodium cacodylate buffer, the EC50 values for DNA condensation were comparable (4 +/- 1 microM) for spermine homologues with n = 4-8, whereas the lower and higher homologues provoked DNA condensation at higher EC50 values. The EC50 values increased with an increase in the monovalent ion (Na+) concentration in the buffer. The slope of a plot of log [EC50(polyamine4+)] against log [Na+] was approximately 1.5 for polyamines with even number values of n, whereas the slope value was approximately 1 for compounds with odd number values of n. Dynamic light scattering measurements showed the presence of compact particles with hydrodynamic radii (Rh) of about 40-50 nm for compounds with n = 3-6. Rh increased with further increase in methylene chain length separating the secondary amino groups of the polyamines (Rh = 60-70 nm for n = 7-10 and >100 nm for n = 11 and 12). Determination of the relative binding affinity of polyamines to DNA using an ethidium bromide displacement assay showed that homologues with n = 2 and 3 as well as those with n > 7 had significantly lower DNA binding affinity compared to spermine and homologues with n = 5 and 6. These data suggest that the chemical structure of isovalent polyamines exerts a profound influence on their ability to recognize and condense DNA, and on the size of the DNA condensates formed in aqueous solution.  相似文献   

17.
The temperature dependencies of the infrared spectra of pure and cholesterol-containing multibilayers of 1,2-dipalmitoyl-sn-glycero-3-phosphocholine were studied using Fourier transform infrared techniques. A comparison of the spectroscopic data showed the retention of a melting phenomenon at 60 mol% cholesterol content, and the retention of some all-trans conformations in the liquid-crystalline phase. It is also demonstrated that at temperatures less than 30 degrees C, the cholesterol-containing 1,2-dipalmitoyl-sn-glycero-3-phosphocholine multibilayers still contain a small amount of pure 1,2-dipalmitoyl-sn-glycero-3-phosphocholine, packed in an orthorhombic subcell lattice. Spectral changes were found in the absorptions characteristic of the phospholipid head groups. The addition of cholesterol results in changes in the ester bands, and demonstrates the induction by cholesterol of non-equivalent ester conformations.  相似文献   

18.
S B Dev  L Walters 《Biopolymers》1990,29(1):289-299
To better understand the structural basis of protein-DNA interactions, the conformational changes that accompany these interactions need to be described. In order to develop a methodological approach to this problem, Fourier transform infrared spectroscopy (FTIR) with derivative resolution enhancement has been used to identify conformational changes that occur when a 29-residue synthetic peptide binds nonspecifically to heterogeneous cellular DNA in aqueous solution. The peptide sequence was chosen de novo, in order to rationally design a peptide model that would allow the relationship between DNA binding and the stability of protein secondary structure to be studied. Peptide at a concentration of 100-200 microM produces 50% saturation of heterogeneous phage DNA sequences as well as of short synthetic oligonucleotides. FTIR spectra reveal significant changes in peptide and DNA upon binding. Second-derivative spectra resolve the amide I band of native peptide into components located at 1627 (beta-strand), 1658 (alpha-helix), and 1681 (turn or beta-strand) cm-1, with a distinct shoulder at 1647 cm-1 (disordered structure). Assignment of the 1681 cm-1 vibration to a turn conformation is supported by uv CD studies, which indicate significant amounts of turn structure in unbound peptide. Ultraviolet CD also confirms the existence of disordered and beta-strand regions in the free peptide. Upon interacting with DNA the band at 1681 cm-1 (turn) is no longer seen; a new band appears at 1675 cm-1; the 1627 cm-1 band (beta-strand) is considerably reduced in intensity; the position of the alpha-helical (1658 cm-1) component remains unchanged; the shoulder at 1647 cm-1 (disorder) disappears. The new vibration at 1675 cm-1 is characteristic of beta-strand structures. The asymmetric stretch (vAS) of the DNA phosphates shifts from 1223 (unbound) to 1229 cm-1 (bound); the relative intensities of vAS and the PO2- symmetric stretch (vS) are altered upon peptide binding.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

19.
Fourier transform infrared study of the halorhodopsin chloride pump   总被引:2,自引:0,他引:2  
Halorhodopsin (hR) is a light-driven chloride pump located in the cell membrane of Halobacterium halobium. Fourier transform infrared difference spectroscopy has been used to study structural alterations occurring during the hR photocycle. The frequencies of peaks attributed to the retinylidene chromophore are similar to those observed in the spectra of the related protein bacteriorhodopsin (bR), indicating that in hR as in bR an all-trans----13-cis isomerization occurs during formation of the early bathoproduct. Spectral features due to protein structural alterations are also similar for the bR and hR photocycles. For example, formation of the red-shifted primary photoproducts of both hR and bR results in similar carboxyl peaks in the 1730-1745-cm-1 region. However, in contrast to bR, no further changes are observed in the carboxyl region during subsequent steps in the hR photocycle, indicating that additional carboxyl groups are not directly involved in chloride translocation. Overall, the close similarity of vibrations in hR and bR photoproduct difference spectra supports the existence of some common elements in the molecular mechanisms of energy transduction and active transport by these two proteins.  相似文献   

20.
In previous work (Zurdo J, Fernández-Cabrera C and Ramírez JM (1993) Biochem J 290: 531–537), it had been shown that selective extraction of the carotenoid from the light-harvesting protein 2 (LH2) of Rhodobacter capsulatus induced the dissociation of 800-nm absorbing bacteriochlorophyll (Bchl), a 10-nm red shift of 854-nm Bchl, and a decrease of the stability of the protein in detergent solution. In the present study, the Fourier transform Raman and near-infrared circular dichroism spectra of native and carotenoid-depleted LH2 membrane preparations were compared. It was found that while the coupled carbonyls of 854-nm Bchl remained specifically H-bonded to the peptides after carotenoid extraction, the optical activity of the near-infrared electronic transition was significantly altered. Given the excitonic origin of such optical activity, our data suggest that carotenoid extraction elicits a rearrengement of the chromophore cluster and of the associated polypeptide subunits. This implies a significant role of the carotenoid in maintaining the native quaternary structure of the protein, which would be consistent with the observed dissociation of 800-nm Bchl and the loss of solubilized LH2 stability that result from carotenoid removal. There is no evidence for a similar role of the carotenoid in the LH1 protein.Abbreviations Bchl bacteriochlorophyll - FT Fourier transform - CD circular dichroism - LH1 and LH2 the bacterial light-harvesting proteins 1 and 2 In memoriam of Daniel I. Arnon.  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