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.
Fluorescence depolarization experiments performed on labaled poly-L -proline Forme II suggest the occurrence of aggrgation in water while 6M guanidinium-HCl induces dissociation. The solvent 4M CaCl2 results in a reduction of polymer structural orgganization. These findings corroborate suggestion of polyproline aggregation and solution behavior in aqueous neutral salt solytion (see preceding article).  相似文献   

2.
J Bello 《Biopolymers》1988,27(10):1627-1640
Poly(trimethyl-L-lysine), [Lys(Me3)]n, is converted from random coil to α-helix at about 1/30 of the NaClO4 concentration required by poly(L-lysine), (Lys)n. NaClO4 generates turbidity in [Lys(Me)3]n at concentrations above that required for helix formation, and decreases turbidity above lM NaClO4. The turbidity runs parallel to enhanced, and then decreased, fluorescence of a dansyl label. Helix formation per se does not induce enhanced fluorescence. Increasing NaClO4 concentration increases Tm linearly with log[NaClO4] for both (Lys)n and [Lys(Me3)]n until the denaturing effect of high NaClO4 sets in. Increasing NaClO4 also increases the breadth of the transition. Heating helical [Lys(Me3)]n or (Lys)n does not produce a CD spectrum resembling that of “random-coil” (Lys)n, except for [Lys(Me3)]n at relatively low NaClO4 concentration.  相似文献   

3.
Complex formation between poly(U) and adenosine in solutions of salts that stabilize (Na2SO4), destabilize (NaClO4), or have little effect on the water structure (NaCl), as well as the poly(U)·poly(A) interaction in NaClO4, was studied by equilibrium dialysis and uv spectroscopy. At 3°C and neutral pH, Ado·2 poly(U) is formed in 1M NaCl and 0.33M Na2SO4. In NaClO4 solutions under the same conditions, an Ado·poly(U) was found over the whole range of salt concentration investigated (10 mM?1M), which has not been previously observed under any conditions. The Ado-poly(U) was also found in a NaCl/NaClO4 mixture, the transition from the triple- to the double-helical complex occurring within a narrow range of concentration of added NaClO4. In the presence of 1M NaCl this transition is observed on adding as little as 10 mM NaClO4, i.e., at a [ClO]/[Cl?] ratio of about 1:100. However, when NaClO4 is added to a 1M solution of the stabilizing salt Na2SO4, no transition occurs even at a [ClO]/[SO] ratio of 1:4. Investigation of melting curves and uv spectra has shown that in an equimolar mixture of the polynucleotides, only a double-helical poly(U)·poly(A) exists in 1M NaClO4 at low temperatures; this also holds for 1M NaCl. This changes to a triple-helical 2 poly(U)·poly(A) and then dissociates as the temperature increases. At low temperatures and the poly(U)/poly(A) concentration ratio of 2:1, a mixture of 2 poly(U)·poly(A) and poly(U)·poly(A) was observed in 1M NaClO4, in contrast to the case of 1M NaCl. Thus, sodium perchlorate, a strong destabilizer of water structure, promotes formation of double-helical complexes both in the polynucleotide–monomer and the polynucleotide–polynucleotide systems. Beginning with a sufficiently high ionic strength (μ ? 0.9), a further increase in the salt molarity results in an increase of the poly(U)·adenosine melting temperature in both stabilizing and neutral salts and a decrease in the destabilizing salt. In Na2SO4 concentrations higher than 1.2M Ado·2 poly(U) precipitates at room temperature. Analysis of the binding isotherms and melting profiles of the complexes between poly(U) and adenosine according to Hill's model shows that the cooperativity of binding, due to adenosine stacking on poly(U), increases in the order NaClO4 < NaCl < Na2SO4. The free energy of adenosine stacking on the template is similar to that of hydrogen bonding between adenosine and poly(U) and ranges from ?1 to ?2 kcal/mol. The values of ΔHt [the effective enthalpy of adenosine binding to poly(U) next to an occupied site, obtained from the relationship between complex melting temperature and free monomer concentration at the midpoint of the transition] are ?14.2, ?18.3, and ?16.8 kcal/mol for 1M solutions of NaClO4, NaCl, and Na2SO4, respectively. The results indicate that the effects of anions of the salts studied are related to water structure alterations rather than to their direct interaction with the complexes between poly(U) and adenosine.  相似文献   

4.
The kinetics of the coil-to-helix transition of (dG-dC)3 in M NaCl, 45 mM sodium cacodylate, pH 7, were measured in H2O, D2O, 10 mol % ethanol, 10 mol % urea, and 10 mol % glycerol. At 43°C in H2O the recombination rate is 1.3 ± 0.2 × 107 M?1 s?1; the dissociation rate is 68 ± 10 s?1. The destabilization of the helix in 10 mol % ethanol and 10 mol % urea relative to water is primarily due to a large increase in the helix-dissociation rate. In 10 mol % glycerol, the destabilization of the helix is due to a decrease in the recombination rate and an increase in the dissociation rate. Above 20°C, two exponential decays longer than 1 μs are observed after a temperature jump. The slower relaxation time is 4–10 times faster than the bimolecular component and is independent of oligomer concentration. We attribute this relaxation to a rapid equilibrium between two helical states. At low temperatures and oligomer concentrations of 1 mM or greater, the helices aggregate in 1M NaCl. Experimental data are presented under conditions where aggregation is unimportant and evidence is given that the ΔH-determined spectroscopically is unaffected by aggregation.  相似文献   

5.
Exposure of yeast 80 S ribosomes to chaotropic salts such as NaClO4 or NaSCN at concentrations as low as 0.4 M resulted in complete dissociation and subsequent aggregation of the ribosomal proteins. However, under similar conditions, both NaCl and NaBr did not cause dissociation and aggregation. The protein precipitate obtained by exposing the ribosomes to 0.5 M NaClO4 was free of any rRNA contamination as judged by ultraviolet-absorption analysis. Comparison of the two-dimensional polyacrylamide gel electrophoretic analysis of the above ribosomal protein precipitate with that ribosomal proteins isolated by the standard acetic acid extraction procedure revealed that the protein precipitate contained all the ribosomal proteins. Based on these results, a simple method for the isolation of total ribosomal proteins and rRNA under mild, nondenaturing conditions is proposed. A possible mechanism for the dissociation of proteins from the ribosome by chaotropic salts is also discussed.  相似文献   

6.
1H- and 13C-nmr studies of conformational transitions of random amino acid copolymers containing aromatic residues (Lys50Tyr50)n and (Lys50Phe50)n in the presence of neutral salts were performed to serve as models of the aggregation behavior of polypeptides of biological significance. The 1H and 13C signal intensities of Tyr and Phe residues decreased preferentially with increasing concentration of neutral salts such as NaCl and NaClO4. This behavior contrasts with that of (Lys)n in the presence of similar neutral salts, where the displacement of the 13C signal is clearly seen on transition from the random-coil to the helical conformation. On the basis of the previous conformational studies, the loss of the peak areas is ascribed to the presence of immobilized helical segments by hydrophobic interaction between aromatic side chains. The remaining resonances are due to the residual random-coil regions, since the values of nuclear Overhauser enhancements and chemical shifts are unchanged in the presence and absence of the neutral salts.  相似文献   

7.
M. Kodama  H. Noda  T. Kamata 《Biopolymers》1978,17(4):985-1002
The conformation of amylose in aqueous solution has been found to be dependent on its molecular weight. When the molecular weight of amylose is outside of the so-called “dissolving gap” described by Burchard (6500<Mr<160,000) it behaves as a random coil, whereas when its molecular weight is within the “dissolving gap,” it easily aggregates forming a rigid coil which is the B-type (retrograded) amylose. The conformation of this rigid coil is suggested to be a double helix.  相似文献   

8.
Polyamino acids which are soluble and helical in acetic acid and dichloroacetic acid (DCA) have been observed to undergo a helix to random chain transition upon the addition of lithium salts of strong acids. The transition can be reversed by diluting the salt. Apparently only lithium cations are able to bring about the polycarbobenzoxy-L -lysine (PCBL) transition in acetic acid, whereas the anions display a varying degree of effectiveness; ClO4? > Br? > TSA? > Cl? > NO3?. The lithium salts of carboxylate anions such as OAc? and TFA? do not cause polymer unwinding in acetic acid. Neither do the acids, TSA, HCl, TFA, or DCA induce the transformation in acetic acid. Poly-L -alanine (PLA) in DCA unfolds as LiBr is added, but does not unfold in the presence of 0.5M (CH3)4NBr, 0.25M CsBr, or 0.32M HCl. These results are explained on the basis of a direct interaction of the lithium salt with the polymer amide groups to form an ion-pair complex. The extent to which the union of the ion pair can dissociate from the complex in the low dielectric constant, environment determines the degree of unfolding of the polymer. The anion dissociation equilibrium presumably therefore would lie in the same order as given above. Acids such as HCl and TSA are considered to substantially protonate and ion-pair with the polymer, but do not readily dissociate the anion partner from the complex, and therefore do not produce an unstable positively charged helical structure.  相似文献   

9.
Four oligomers of α-L -glutamic acid with ends blocked by nonionizable protecting groups were studied in 0.1M NaClO4 solution using potentiometric measurements. The titrations were analyzed using a pit-mapping method. The results are compared with those obtained with the corresponding polymer. The successive dissociation constants of the carboxylic side chains are obtained for oligomers with degrees of polymerization varying from 1 to 6. Then the successive stability constants corresponding to the association Cu2+–oligomer are given, and the MpHqLr distribution is calculated for each oligomer introducing MHqL complexes.  相似文献   

10.
A 13C-nmr study of the salt-induced helix–coil transition of the basic polypeptides poly(L -lysine) [(Lys)n], poly(L -arginine) [(Arg)n], and poly (L -ornithine) [(Orn)n] was performed to serve as a reference of the helical portion of histones and other proteins. As is the case with pH-induced helix–coil transition, the downfield displacement of the Cα and carbonyl carbon signals are observed in the helical state. The upfield shift of the Cβ signals, on the other hand, is noted in the salt-induced transition. Regardless of the differences in the side chains and also the salts used, very similar helix-induced chemical shifts are obtained for (Lys)n and (Arg)n. However, the displacement of the Cα, Cβ, and carbonyl carbons of (Orn)n in the presence of 4M NaClO4 is found to be almost 50% of that of (Lys)n and (Arg)n. This is explained by the fact that the maximum helical content is about 50%, consistent with the ORD result. Further, the motion of the backbone and side chains of the helical from was estimated by measuring the spin-lattice relaxation time (T1), nuclear Overhauser enhancement (NOE), and line width. In the case of (Lys)n, the motion of the side chains is charged very little in comparison with that of the random coil. Indicating that the aggregation of the salt-induced helix is small in contrast to that of the pH-induced helix. For (Arg)n, however, the precipitate of the helical polymers is mainly due to aggregation.  相似文献   

11.
The polyproline‐II helix is the most extended naturally occurring helical structure and is widely present in polar, exposed stretches and “unstructured” denatured regions of polypeptides. Can it be hydrophobic? In this study, we address this question using oligomeric peptides formed by a hydrophobic proline analogue, (2S,3aS,7aS)‐octahydroindole‐2‐carboxylic acid (Oic). Previously, we found the molecular principles underlying the structural stability of the polyproline‐II conformation in these oligomers, whereas the hydrophobicity of the peptide constructs remains to be examined. Therefore, we investigated the octan‐1‐ol/water partitioning and inclusion in detergent micelles of the oligo‐Oic peptides. The results showed that the hydrophobicity is remarkably enhanced in longer oligomeric sequences, and the oligo‐Oic peptides with 3 to 4 residues and higher are specific towards hydrophobic environments. This contrasts significantly to the parent oligoproline peptides, which were moderately hydrophilic. With these findings, we have demonstrated that the polyproline‐II structure is compatible with nonpolar media, whereas additional manipulations of the terminal functionalities feature solubility in extremely nonpolar solvents such as hexane.  相似文献   

12.
Laser light-scattering has been used to investigate the size of native proteoglycan aggregates (PGA-aA1) from day-8 chick limb-bud chondrocyte cultures isolated under associative extraction and purification conditions in 0.4M guanidinium chloride (GdnHCl) solution. Dynamic light-scattering measurements yielded a hydrodynamic radius, Rs, of 244 ± 10 nm for PGA-aA1 in 0.4M GdnHCl, and a weight-average molecular weight (M w) of 150 ± 50 × 106 was obtained from a Zimm plot. Disaggregation in 4.0M GdnHCl aqueous solution yielded proteoglycan subunits (PGS) with Rs = 39 ± 2 nm, M w = 1.6 ± 0.3 × 106, which reassembled in 0.4M GdnHCl to form “reconstituted native” aggregates (PGA-raA1) with Rs = 121 ± 6 nm, M w = 17 ± 3 × 106. A second specimen of PGA-aA1 had Rs = 192 ± 10 nm, M w = 100 ± 10 × 106. The latter value was estimated from an empirical relationship between M w and Rs. After dissociation, this specimen reassembled to form PGA-raA1 with Rs = 85 ± 5 nm, M w = 12 ± 1 × 106. These data are compared with those for a specimen of reconstituted aggregate (PGA-A1) that had been extracted under dissociative conditions and then reaggregated by dialysis to 0.4M GdnHCl aqueous solution, for which Rs = 138 ± 9 nm, M w = 45 ± 8 × 106. From these values, we have calculated the weight-average number of subunits per aggregate Nw: 111 for PGA-aA1 and 12 for raA1 (70 and 7 for the second PGA-aA1 and PGA-raA1 specimen, respectively) as compared to 32 for PGA-A1. The numbers of subunits per aggregate were also determined from electron micrographs of spread specimens. The latter results show the same trends as those obtained by light scattering, but lead in each case to lower numbers of subunits per aggregate. These data demonstrate conclusively that PGA samples exhibit a higher degree of aggregation in solution than visualized in typical electron microscopy (EM) preparations, probably due to disaggregation during EM specimen preparation. Since Nw determined both by light scattering (LS) and by EM are larger for native versus reconstituted aggregate samples, our data point to a more compact aggregation of subunits along the hyaluronic acid (HA) chains in the former.  相似文献   

13.
Helix-coil dynamics of a Z-helix hairpin   总被引:1,自引:0,他引:1  
The helix–coil transition of a Z-helix hairpin formed from d(C-G)5T4(C-G)5 has been characterized by equilibrium melting and temperature jump experiments in 5M NaClO4 and 10 mM Na2HPO4, pH 7.0. The melting curve can be represented by a simple all-or-none transition with a midpoint at 81.6 ± 0.4°C and an enthalpy change of 287 ± 15 kJ/mole. The temperature jump relaxation can be described by single exponentials at a reasonable accuracy. Amplitudes measured as a function of temperature provide equilibrium parameters consistent with those derived from equilibrium melting curves. The rate constants of Z-helix formation are found in the range from 1800 s?1 at 70°C to 800 s?1 at 90°C and are associated with an activation enthalpy of ?(50 ± 10) kJ/mole, whereas the rate constants of helix dissociation are found in the range from 200 s?1 at 70°C to 4500 s?1 at 90°C with an activation enthalpy +235 kJ/mole. These parameters are consistent with a requirement of 3–4 base pairs for helix nucleation. Apparently nucleation occurs in the Z-helix conformation, because a separate slow step corresponding to a B to Z transition has not been observed. In summary, the dynamics of the Z-helix–coil transition is very similar to that of previously investigated right-handed double helices.  相似文献   

14.
Poly(Lys(HBr)-Gly-Pro-Pro-Gly-Pro) has been synthesized and studied by circular dichroism (CD) spectroscopy. It is apparently the first polyhexapeptide collagen model reported with an ionizable side chain. The monomer (ε-(p-nitrobenzyloxycarbonyl)-Lys-Gly-Pro-Pro-Gly-Pro-p-nitrophenyl-ester) was prepared by a stepwise strategy employing active esters. Polymerization in N,N-dimethyl formamide, followed by removal of the Lys side chain protection with HBr/acetic acid, gave a polydisperse product. Fractionation was accomplished by gel filtration chromatography. The polydisperse material had a molecular weight (Mr = 5–17,000). High molecular weight fractions from triple helices under concentrated conditions at 2°C. The triple helical structure gives a CD pattern very similar to that of collagen and its triple helical analogs. However, unlike collagen, the polyhexapeptide undergoes spontaneous dissociation at temperatures substantially below the melting temperature from a triple helical form to single chains. This process is promoted at low concentrations, high temperature, neutral pH, and low molecular weight, and is apparently due, in large part, to unfavorable ionic side-chain interactions. In addition to this relatively slow “ionic” dissociation the triple helical polypeptide may be thermally dissociated in a manner similar to collagen. The thermal denaturation is a relatively fast process compared with ionic dissociation. A high molecular weight fraction (3 × Mr = 48,000) was found to melt at 42°C at neutral pH but increased to 54°C at pH 12 where the lysyl side chains are predominantly deprotonated. Furthermore, reconstitution of triple helices appeared to be more readily achieved at high pH. Thus it is concluded that ionic repulsion between side chains causes destabilization of the triple helix and hinders reconstitution.  相似文献   

15.
A semi-empirical conformational energy calculation has been performed on the ionizable polydipeptide poly(Glu-Ala). The results indicate that; (1)the ionized polymer assumes the lefthanded extended helix conformation is aqueous solution; (2) the poly(Glu-Ala) extended helix is less stable than that of polyglutamic acid; (3) the unionized polydipeptide will preferentially assume the “β-helical” conformation (isolated 21 degenerate helix) in aqueous solution. These conclusions are supported by experiment.  相似文献   

16.
The ultraviolet absorption, linear dichroism, circular dichroism, and oriented circular dichroism of collagen are reported and the spectra are resolved into a self-consistent set of bands in accord with exciton theory. The parallel band at 200 nm has 40% of the π → π* intensity; the perpendicular band is placed at 189 nm yielding a splitting of 2700 cm?1. The circular dichroism is resolved into two Gaussians at λ and λτ (rotational strengths +14 × 10?40 and ?32 × 10?40 esu2. cm2) plus a large non-Gaussian (“helix”) band with ampplitude ?25,000° at 201 nm. These data appear to be in reasonably good accord with recent calculations. Measurements of the absorption, linear dichroism and circular dichroism of polyproline I and II are also reported and are resolved into their component bands. Polyproline I is in good accord with exciton theory, whereas polyproline II remains unsatisfactory.  相似文献   

17.
Nucleic acid recognition is often mediated by α‐helices or disordered regions that fold into α‐helix on binding. A peptide bearing the DNA recognition helix of HPV16 E2 displays type II polyproline (PII) structure as judged by pH, temperature, and solvent effects on the CD spectra. NMR experiments indicate that the canonical α‐helix is stabilized at the N‐terminus, while the PII forms at the C‐terminus half of the peptide. Re‐examination of the dihedral angles of the DNA binding helix in the crystal structure and analysis of the NMR chemical shift indexes confirm that the N‐terminus half is a canonical α‐helix, while the C‐terminal half adopts a 310 helix structure. These regions precisely match two locally driven folding nucleii, which partake in the native hydrophobic core and modulate a conformational switch in the DNA binding helix. The peptide shows only weak and unspecific residual DNA binding, 104‐fold lower affinity, and 500‐fold lower discrimination capacity compared with the domain. Thus, the precise side chain conformation required for modulated and tight physiological binding by HPV E2 is largely determined by the noncanonical strained α‐helix conformation, “presented” by this unique architecture. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 432–443, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

18.
The binding of cupric ion (Cu++) to DNA was followed by spectrophotometry, melting profiles, and hydrodynamic techniques, in 0. 1M NaClO4 and at pH 5. 6. A small amount of Cu++ is bound specifically to bases (about 1 Cu++ per 20 nucleotides), in agreement with polarographic and EPR data. A preferential stabilization of G–C pairs and only a slight increase of the flexibility of the molecule were observed. In 5 × 10?3M NaClO4, a higher number of nonhomogeneous binding sites is found by spectrophotometry. It is concluded that at least two types of sites are available for Cu++. The first one, where Cu++ is chelating N7 of purines to phosphate, is observed only at low ionic strength and destabilizes the double helix. The second exists mainly at 0, 1M or higher ionic strength. All the sites are identical and could be attributed to two successive guanine residues in the same strand. Similar behavior was found for other divalent cations, e. g., Fe++, Mn++, and Co++.  相似文献   

19.
Titration of an aqueous solution of sodium poly-L-glutamate with a strong acid usually produces turbidity and precipitation before the equivalence point is reached. In 1M sodium p-toluenesulfonate and 1M sodium methylcyclohexanesulfonate aggregation was delayed for days to months. In very concentrated tetra-n-butylammonium chloride and bromide the polymeric acid dissolved and could be titrated with NaOH. However, potentiometric titration curves in these solutions did not yield information (by way of plots of apparent pK versus the degree of neutralization) about the helix-to-coil transition of the polymer. It was argued in addition that the apparent pK of a weak polyelectrolyte should not be calculated from titrations in concentrated salt solutions since it is a mixed or composite quantity. It contains not only the effect of the salt on the dissociation of the weak electrolyte but its effect on the activity coefficient of hydrogen ions as well. Circular dichroic spectra were therefore measured a t various degrees of neutralization of poly(L -glutamic acid) in a number of aqueous salt solutions and mixtures of organic solvents with water. It was found that the undissociated polymer in the concentrated methylcyclohexanesulfonate and quaternary ammonium halide solutions had the spectrum of a right-handed α helix. The n → π* band at 222 mμ was used as a measure of the fraction of polymer in the helical conformation. The value of ? Δε for the undissociated polyacid in these organic electrolyte solutions was 11.4. By means of this value and a number of assumptions, the fraction of helix (fh) as a function of α, the degree of polyacid neutralization, was calculated for the different solvent mixtures. An empirical equation was used to describe the variation of fh with α, fh = 1/(1 + e?a+bα), in which b represents the degree of cooperativeness of the transition, and a is a measure of the effect of the medium on the onset of the transition. The values of b did not differ very much from one another, suggesting that the cooperativeness of the transition was not sensitive to changes in the medium. On the other hand, the value of a (or its equivalent, the value of α at the halfway point in the transition) was more dependent on the solvent. Comparisons of these results with those of some other workers were made by means of the empirical equation.  相似文献   

20.
Infrared spectroscopy and equilibrium binding studies have been used to determine the combining ratio of adenosine and poly(U) in 1M NaClO4 solution. In contrast to a recent report, we find exclusively 1:2 stoichiometry in this solvent. Ultraviolet and infrared spectroscopy have been employed to show that the stoichiometry of poly(A) interaction with poly(U) is unaffected by 1M NaClO4.  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