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.
Changes in the 31P-nmr spectra of sonicated natural DNA fragments were investigated in ethanol solutions where the fragments underwent, as checked by CD, the B-to-A conformational transition. The study produced the following conclusions: (1) The high DNA concentrations used for the 31P-nmr measurements promote the transition compared to dilute solutions that are commonly used for CD measurements. (2) The B-to-A transition was reflected in a cooperative downfield shift of the DNA 31P-nmr resonance, consistent with unwinding of the double helix. (3) Prior to the transition, the changes in chemical shift of double-and single-stranded DNAs were almost identical. It thus appears that the effect of ethanol on the geometry and hydration of phosphodiester linkages does not depend heavily on DNA base–base interactions. (4) The A-form resonances were 30–40% narrower than the B-form resonances, which is attributed to marked sequence-dependent variations in the latter conformation and to their reduction in the former. (5) The B-form DNA aggregated in the concentrated 31P-nmr samples in the presence of ethanol, judged from a milky opalescence of the solution and a substantial broadening of its 31P-nmr resonance. The broadening abruptly disappeared as soon as DNA adopted the A-form so that DNA, in dependence on the secondary structure, showed different tendencies to condense in the presence of ethanol. The condensation increased cooperativity of the B-to-A interconversion.  相似文献   

2.
S P Fodor  P A Starr  T G Spiro 《Biopolymers》1985,24(8):1493-1500
Raman spectra have been recorded for poly(dG-dT) · poly(dA-dC) and poly(dA-dT) · poly(dA-dT) in low salt and at high concentrations of CsF. Poly(dG-dT) · poly(dA-dC) shows no change in the 682-cm?1 guanine mode, demonstrating the absence of the Z-structure at high salt. The 790-cm?1 phosphodiester symmetric stretch, however, shifts up 5 cm?1 in 4.3M CsF, suggesting a slight conformational change, associated with ion binding or hydration changes. Poly(dA-dT) · poly(dA-dT) shows an additional broad band at 816 cm?1, attributed to the phosphodiester modes associated with the C3′-endo deoxyribose units in the alternating B-structure. In this case, both the 841- and the 816-cm?1 asymmetric phosphodiester stretches, associated with the C2′- and C3′-endo units, shift down on addition of CsF in a sequential manner. Correlation of this sequence with that previously observed for the two 31P-nmr resonances, establishes that the phosphodiester stretching frequencies depend on the conformation of the 5′-sugar, and not on the 3′-sugar.  相似文献   

3.
We examined three sonicated, specific-seqiemce polydeoxynucleotides in solution over a wide range of concentrations of several salts by 13P-nmr spectroscopy, and we found that the alternating copolymer poly(dAdT)·poly(dAdT) exhibits a dinucleotide repeat unit in all five salts and at all concentrations studied, as indicated by the presence of a doubled in its 31P-nmr spectra. The two components of the doublet show selective shift effects. The upfield component is assigned to dApdT in the gauche?-gauche? conformation and shifts upfield in all four monovalent salts used, relative to a single-stranded oligonucleotide control. The downfield component is assigned to dTpdA in the trans-gauche? conformation and shifts downfield with increasing CsF concentration but remains essentially constant in LiCl, NaCl, and CsCl. These changes indicate a fast noncooperative transition for poly(dAdT)·poly-(dAdT) from a presumed right-handed dinucleotide-repeat B-form to another conformation with a dinucleotide-repeat structure, via a continuum of structures that may differ in the extent of the winding of the double helix. Ethanol causes the upfield component to collapse into the other component, indicating conversion to a structure with a mononucleotide repeat unit and a trans-gauche? conformation. Up to 1M Mg2+ appears to have no significant effect on the phosphodiester conformations of poly(dAdT)·poly(dAdT). By contrast, poly-(dGdC)·poly(dGdC) gives a slow cooperative transition from what is considered to be a right-handed regular B-form to a left-handed Z-form on increasing MgCl2 and NaCl concentrations, although we observed no changes in chemical shifts below the transition points. The homopolymer poly(dA)·poly(dT) exhibits no unusual shift effects or transitions upon the addition of salts when compared to the oligonucleotide control and is considered to be a regular B-form with a gauche?-gauche? phosphodiester backbone conformation. These differences emphasize the distinct secondary structures of DNAs of different sequences and their selective responses to changes in solution conditions.  相似文献   

4.
31P-nmr has been used to investigate the specific interaction of three divalent metal ions, Mg2+, Mn2+, and Co+2, with the phosphate groups of DNA. Mg2+ is found to have no significant effect on any of the 31P-nmr parameters (chemical shift, line-width, T1, T2, and NOE) over a concentration range extending from 20 to 160 mM. The two paramagnetic ions, Mn2+ and Co2+, on the other hand, significantly change the 31P relaxation rates even at very low levels. From an analysis of the paramagnetic contributions to the spin–lattice and spin–spin relaxation rates, the effective internuclear metal–phosphorus distances are found to be 4.5 ± 0.5 and 4.1 ± 0.5 Å for Mn2+ and Co2+, respectively, corresponding to only 15 ± 5% of the total bound Mn2+ and Co2+ being directly coordinated to the phosphate groups (inner-sphere complexes). This result is independent of any assumptions regarding the location of the remaining metal ions which may be bound either as outer-sphere complexes relative to the phosphate groups or elsewhere on the DNA, possibly to the bases. Studies of the temperature effects on the 31P relaxation rates of DNA in the absence and presence of Mn2+ and Co2+ yielded kinetic and thermodynamic parameters which characterize the association and dissociation of the metal ions from the phosphate groups. A two-step model was used in the analysis of the kinetic data. The lifetimes of the inner-sphere complexes are 3 × 10?7 and 1.4 × 10?5 s for Mn2+ and Co2+, respectively. The rates of formation of the inner-sphere complexes with the phosphate are found to be about two orders of magnitude slower than the rate of the exchange of the water of hydration of the metal ions, suggesting that expulsion of water is not the rate-determining step in the formation of the inner-sphere complexes. Competition experiments demonstrate that the binding of Mg2+ ions is 3–4 times weaker than the binding of either Mn2+ or Co2+. Since the contribution from direct phosphate coordination to the total binding strength of these metal ion complexes is small (~15%), the higher binding strength of Mn2+ and Co2+ may be attributed either to base binding or to formation of stronger outer-sphere metal–phosphate complexes. At high levels of divalent metal ions, and when the metal ion concentration exceeds the DNA–phosphate concentration, the fraction of inner-sphere phosphate binding increases. In the presence of very high levels of Mg2+ (e.g., 3.1M), the inner-sphere ? outer-sphere equilibrium is shifted toward ~100% inner-sphere binding. A comparison of our DNA results and previous results obtained with tRNA indicates that tRNA and DNA have very similar divalent metal ion binding properties. A comparison of the present results with the predictions of polyelectrolyte theories is presented.  相似文献   

5.
C. P. Beetz  G. Ascarelli 《Biopolymers》1982,21(8):1569-1586
We have measured the ir absorption of 5′CMP, 5′IMP, and poly(I)·poly(C) from ~25 to ~500 cm?1. From a comparison of the data with the previously measured absorption of the corresponding nucleosides and bases we can identify several “lines” associated with the deformation of the ribose ring. Out-of-plane deformation of the bases contributes strongly to vibrations near 200 cm?1. The same ribose vibrations observed in the nucleotides are found in poly(I)·poly(C). They sharpen with increasing water absorption. A study of the spectra of poly(I)·poly(C) as a function of the adsorbed water indicates that water does not contribute in a purely additive fashion to the polynucleotide spectrum but depends on the conformation of the helix. However, the only spectral feature that shifts drastically with conformation is near 45 cm?1. Measurements at cryogenic temperatures indicate some sharpening of the spectrum of poly(I)·poly(C). Instead, no sharpening is observed in the spectrum of the nucleotides. Shear degradation of poly(I)·poly(C) produces significant spectral changes in the 200-cm?1 region and sharpening of the features assigned to the low-frequency ribose-ring vibrations.  相似文献   

6.
Using direct difference ir and laser Raman spectroscopy, the sequential hydration of hen egg-white lysozyme was monitored. The ir data allowed us to identify some specific molecular hydration events that occur as water is added, whereas the Raman is interpreted in terms of conformational changes. The largest of these solvent-induced changes occurs below the hydration level at which activity commences.  相似文献   

7.
We report proton magnetic resonance studies of a series of lysine oligopeptides in H2O solution. At pH 5 the protonated ε-amino groups are seen as broad resonances; the peptide NH proton resonances are split by spin–spin coupling with the Cα-H proton, and appear at positions which depend on position in the chain and on chain length. Assignments were made by the europium shift method, and we observed the expected effect of catalysis by the terminal —NH3+ of exchange of the adjacent peptide NH. Coupling constants and the temperature coefficient of chemical shift values were consistent with a non-hydrogen-bonded structure for the oligolysines. The rate and mechanism of NH hydrogen exchange were investigated by line-broadening measurements of the peptide protons as a function of pH. Exchange was found to be OH? catalyzed, with large differences in the rate depending on position in the chain. Preliminary studies of the complex between double-helical d(pA)3pGpC(pT)3 and tetra(L -lysine) were performed using 1H- and 31P-nmr techniques. Pmr spectra of the complex at pH values ranging from 3.98 to 8.15 showed very complicated patterns. Downfield shifts and reduction in exchange rates were observed for several tetra(L -lysine) protons. 31P-nmr spectra of the complex reveal an upfield shift of 1 ppm for 3′-5′ phosphate diester resonances on complexation. 31P T1 relaxation times change little on complex formation at low temperature but are altered at higher temperature.  相似文献   

8.
The interaction of propidium with three self-complementary oligodeoxyribonucleotides has been investigated by 1H- (base-pair imino proton assigned by 1D NOE and saturation transfer methods) and 31P-nmr as a function of ratio of propidium to oligomer (from zero to saturation) and temperature. The three oligomers are dTATATGCGCATATA (1), dTATATGTGCATATA (2), which has the same sequence as 1 except for the mismatched base pair at position 7, and dTATGTGCATA (3), which is a shortened version of 2. The imino proton chemical-shift changes of 1 on titration with propidium can be explained by the effects of the ring-current anisotropy of propidium at intercalation (3.4 Å) and next-neighbor sites (6.8 Å). The results indicate that propidium binds with neighbor exclusion but with no significant specificity for any intercalation site in the sequence of 1. The addition of propidium to 1 results in general downfield shifts of all 31P signals, as expected for a nonspecific intercalator. Imino and 31P-nmr spectra for 2 indicate that this oligomer forms a hydrogen-bonded G · T base pair at position 7 with little change in base pairing and stacking of base pairs 1–6 compared to 1. The results for addition of propidium to 2 and 3 are quite different than with 1. At low ratio only secondary shifts (6.8 Å) are seen for the G and T imino protons of base-pair 7 on addition of propidium. At higher ratios of propidium, the signals for these G and T protons are lost in 2 and severely broadened in 3, even at low temperature. The other potential intercalation sites in 2 and 3 appear to bind propidium strongly and without significant specificity as with 1. 31P spectra of 2 in the presence of propidium show the expected downfield shifts and broadening. Thus, the minor differences in local helix geometry in 1, and in 2 and 3, away from the G · T base pair do not significantly affect propidium intercalation specificity. Having one or two G · T base pairs at a site, however, makes intercalation in the standard manner significantly less favorable.  相似文献   

9.
Interactions of meso-tetra(4-N-methylpyridyl)porphyrin [TMpyP(4)], meso-tetra(2-N-methylpyridyl)porphyrin [TMpyP(2)], and meso-tetra(para-N-trimethylanilinium)porphyrin (TMAP) with several native and synthetic DNAs were studied by a variety of physical techniques: nmr (31P and 1H), absorption spectroscopy, viscosity, and flow dichroism (FD). Of the three porphyrins studied, only the interaction of TMpyP(4) with poly [d(G-C)2] was fully consistent with intercalation. In particular, a large increase in viscosity, a downfield 31P-nmr signal (ca. -1 ppm), and upfield imino proton signals (11 to 12 ppm range) were observed. Comparison of the effects of TMpyP(4) on DNAs of different GC contents revealed larger changes in solution viscosity with increased GC content. However, the characteristic changes in 31P- and 1H-nmr spectra were not observed. The viscosity increases observed in studies with poly[d(A-C)(G-T)] and C. Perf. DNA were much lower than with poly[d(G-C)2], M. Lys. DNA, and calf thymus DNA. Thus, GC sequence and content are clearly important. The principal change in the 31P-nmr signal of native DNA is the appearance of a very broad shoulder centered at ca. -2.0 ppm, which is larger in M. Lys. DNA than in C. Perf. DNA. FD studies indicate highly ordered TMpyP(4) cations arranged perpendicular to the DNA axis of calf thymus DNA. Together, these results suggest the major effects of TMpyP(4) on DNA properties are due to strong GC-binding interactions that influence DNA structure. The data are consistent with combined intercalative and outside binding interactions of TMpyP(4) with GC regions of DNA. In contrast, similar studies with TMAP suggest that it influences AT regions of DNA by an outside binding mode. On the other hand, TMpyP(2) effects on DNA properties are consistent with nonselective outside binding.  相似文献   

10.
Bisintercalation of ditercalinium, a potent antitumoral 7H-pyriodo[4,3-c]carbazole rigid dimer, into the self-complementary tetranucleotides d(CpGpCpG)2, d(m5CpGpm5CpG) and the self-complementary hexanucleotide d(CpGpApTpCpG)2 was investigated by 162-MHz 31P-nmr. The slow exchange, on the nmr time scale, observed between the free and complexed nucleotides allows identification of the phosphorus signals in the complexes through two-dimensional chemical exchange spectroscopy. Differences in 31P chemical shifts upon intercalation are discussed in relation to the complex geometry and nature of the drug.  相似文献   

11.
We studied films of poly(L -tyrosine) with hydrogen phosphate (residue/phosphate, 1:1) by ir spectroscopy. The influences of the alkali cations (Li+, Na+, K+) and of the degree of hydration were clarified. If Li+ ions are present, the OH ??OP hydrogen bonds formed in the dried films between the tyrosine OH groups and hydrogen phosphate are asymmetrical. The formation of hydrogen phosphate–hydrogen phosphate hydrogen bonds is prevented by the presence of the Li+ ions. With an increase in the degree of hydration, the tyrosine–phosphate bonds are not broken but become slightly stronger. Completely different behaviour is found if K+ ions are present. In dry films, the OH ??OP ? O? ?HOP hydrogen bonds formed between tyrosine and hydrogen phosphate show large proton polarizability. The tyrosine proton has a noticeable residence time at the acceptor O atom of the phosphate. The difference in the behaviour of the system with K+ ions when compared to the system with Li+ ions can be explained, since the hydrogen acceptor O atom of phosphate ions is more negatively charged due to the weaker influence of the K+ ions. Furthermore, POH ??OP hydrogen bonds between hydrogen phosphate molecules are formed. With an increase in the degree of hydration, the tyrosine–hydrogen phosphate hydrogen bonds are broken, all tyrosine protons are found at the tyrosine residues, and the -PO3? groupings are in a symmetrical environment, indicating that the K+ ions are removed from these groupings. If the degree of hydration increases further, hydrogen-bonded systems such as hydrogen phosphate–water–hydrogen phosphate are formed that show large proton polarizability due to collective proton motion. When Na+ ions are present, the OH ??OP ? O? ?HOP hydrogen bonds formed in dry films still show proton polarizability, but the residence time of the tyrosine proton at the phosphate is very short.  相似文献   

12.
U Burget  G Zundel 《Biopolymers》1987,26(1):95-108
(L -His)n- dihydrogen phosphate systems are studied by ir spectroscopy in the presence of various cations and as a function of the degree of hydration. Ir continua indicate that (I) OH … N ? O?…H+N (IIR) hydrogen bonds are formed and that these bonds show high proton polarizability, which increases from the Li+ to the K+ system. In the K+?system, His-Pi-Pi chains are formed, showing particularly high proton polarizability due to collective proton motion within both hydrogen bonds. The OH N ? O?…H?N equilibria are determined from ir bands. With the Li+ system, 55% of the protons are present at the histidine residues; this percentage is smaller with the Na+ system (41%), and amounts to only 32% with the K+ system. With the increasing degree of hydration the removal of the degeneracy of νas?PO2?3 vanishes, indicating loosening of the cations from the phosphates. Nevertheless, the hydrogen bond acceptor O atom becomes more negative; a shift of the equilibrium to the right is observed in the OH… N ? O?…H+N bond. This is explained by the strong interaction of the dipole of the hydrogen bonds with the water molecules. All these results show that protons can be shifted easily in these hydrogen bonds due to their high proton polarizability. The transfer equilibria can be controlled easily by local electrical fields. In addition, these results may be of significance when phosphates interact with proteins.  相似文献   

13.
Analysis of the proton-decoupled 31P-nuclear magnetic resonance (NMR) spectrum of fully hydrated Typha latifolia pollen revealed the presence of two main peaks: A broad asymmetrical component of a `bilayer' lineshape and a much narrower symmetrical component originating from phosphorus compounds undergoing rapid isotropic motion. From (a) 31P-NMR experiments on the hydrated total pollen phospholipids, (b) saturation transfer 31P-NMR experiments, and (c) the fraction of lipid phosphate in the pollen, it can be concluded that the great majority of the endogenous phospholipids are arranged in extended bilayers in which the lipid phosphates undergo fast (τc < 10−6 second) long axis rotation. This bilayer arrangement of phospholipids was observed in the pollen down to hydration levels of at least 10.9% moisture content. At the lowest level of pollen hydration examined (5.2%) the 31P-NMR spectrum had a solid state lineshape demonstrating that all the phosphorus-containing compounds (including the phospholipids) were virtually immobile.  相似文献   

14.
Methanosarcina frisia accumulates phosphate up to 14% of its dry weight. The phosphate is stored as long-chain polyphosphates as shown by 31P-NMR investigations. Further results show that the accumulation of phosphates is substrate-dependent. In the presence of H2 and CO2 as the only carbon and energy source 180 mg of PO inf4 sup3- /g protein were accumulated, whereas 260 mg PO inf4 sup3- /g protein were accumulated in the presence of methanol. This is far more than necessary for the maintenance of essential metabolic pathways. In addition, the 31P-NMR studies show the occurrence of cyclic 2,3-diphosphoglycerate in Methanosarcina frisia. The role of the phosphate metabolites in cell metabolism are discussed.Abbreviations M. Methanosarcina - CCP cyclic 2,3-diphosphoglycerate  相似文献   

15.
1. Shifts in the 1H and 31P-nmr signals originating from the outer and inner phosphorylcholine head-groups and from the lipid acyl chains are observed when phosphatidylcholine vesicles are treated with increasing extravesicular concentrations of the lanthanides Eu3+, Pr3+, Yb3+, and Dy3+. 2. The addition of KNCS to increase the binding of the lanthanide ions to the outer head-groups is used to demonstrate that the intravesicular group shifts are not caused by bulk susceptibility effects. 3. The magnitude and direction of the observed shifts in the 1H-nmr spectrum are shown to be consistent with (a) pseudocontact interaction of the paramagnetic lanthanide ions with the outer phospholipid head-groups, (b) current views of the conformation of the phosphatidylcholine head-group in the presence of lanthanides, and (c) a conservation of magnetic field within the vesicles due to their spherical nature. 4. Variation of the shifts with temperature are compared for egg phosphatidylcholine and dipalmitoyl phosphatidylcholine. The temperature variation in shifts is also used to study phase transitions in each monolayer and phase separations in mixed lipid systems.  相似文献   

16.
The temperature dependence of the 31P NMR spectra of the ethidium complexes with poly(A) X oligo(U) and the 31P spectra of phenylalanine tRNA (yeast) in various molar ratios of ethidium ion (Et) are presented. In the poly(A) X oligo(U) X Et complex, a new peak about 2.0 ppm downfield from the double-helix peak appears. We have assigned this peak to phosphates perturbed by ethidium. The chemical shift of this peak is consistent with the intercalation mode of binding and provides additional support for our hypothesis that 31P shifts are sensitive probes of phosphate ester conformations. The main effect of ethidium on the 31P spectra of tRNAPhe is the broadening of several of the scattered signals. These scattered signals are associated with phosphates involved in tertiary interactions. We propose that these broadened signals arise from phosphates near the Et binding site.  相似文献   

17.
31P- and 1H-nmr and laser Raman spectra have been obtained for poly[d(G-T)]·[d(C-A)] and poly[d(A-T)] as a function of both temperature and salt. The 31P spectrum of poly[d(G-T)]·[d(C-A)] appears as a quadruplet whose resonances undergo separation upon addition of CsCl to 5.5M. 1H-nmr measurements are assigned and reported as a function of temperature and CsCl concentration. One dimensional nuclear Overhauser effect (NOE) difference spectra are also reported for poly[d(G-T)]·[d(C-A)] at low salt. NOE enhancements between the H8 protons of the purines and the C5 protons of the pyrimidines, (H and CH3) and between the base and H-2′,2″ protons indicate a right-handed B-DNA conformation for this polymer. The NOE patterns for the TH3 and GH1 protons in H2O indicate a Watson–Crick hydrogen-bonding scheme. At high CsCl concentrations there are upfield shifts for selected sugar protons and the AH2 proton. In addition, laser Raman spectra for poly[d(A-T)] and poly[d(G-T)]·[d(C-A)] indicate B-type conformations in low and high CsCl, with predominantly C2′-endo sugar conformations for both polymers. Also, changes in base-ring vibrations indicate that Cs+ binds to O2 of thymine and possibly N3 of adenine in poly[d(G-T)]·[d(C-A)] but not in poly[d(A-T)]. Further, 1H measurements are reported for poly[d(A-T)] as a function of temperature in high CsCl concentrations. On going to high CsCl there are selective upfield shifts, with the most dramatic being observed for TH1′. At high temperature some of the protons undergo severe changes in linewidths. Those protons that undergo the largest upfield shifts also undergo the most dramatic changes in linewidths. In particular TH1′, TCH3, AH1′, AH2, and TH6 all undergo large changes in linewidths, whereas AH8 and all the H-2′,2″ protons remain essentially constant. The maximum linewidth occurs at the same temperature for all protons (65°C). This transition does not occur for d(G-T)·d(C-A) at 65°C or at any other temperature studied. These changes are cooperative in nature and can be rationalized as a temperature-induced equilibrium between bound and unbound Cs+, with duplex and single-stranded DNA. NOE measurements for poly[d(A-T)] indicate that at high Cs+ the polymer is in a right-handed B-conformation. Assignments and NOE effects for the low-salt 1H spectra of poly[d(A-T)] agree with those of Assa-Munt and Kearns [(1984) Biochemistry 23 , 791–796] and provide a basis for analysis of the high Cs+ spectra. These results indicate that both polymers adopt a B-type conformation in both low and high salt. However, a significant variation is the ability of the phosphate backbone to adopt a repeat dependent upon the base sequence. This feature is common to poly[d(G-T)]·[d(C-A)], poly[d(A-T)], and some other pyr–pur polymers [J. S. Cohen, J. B. Wouten & C. L Chatterjee (1981) Biochemistry 20 , 3049–3055] but not poly[d(G-C)].  相似文献   

18.
The biological and chemical basis of vanadium action in fungi is relatively poorly understood. In the present study, we investigate the influence of vanadate (V5+) on phosphate metabolism of Phycomyces blakesleeanus. Addition of V5+ caused increase of sugar phosphates signal intensities in 31P NMR spectra in vivo. HPLC analysis of mycelial phosphate extracts demonstrated increased concentrations of glucose 6 phosphate, fructose 6 phosphate, fructose 1, 6 phosphate and glucose 1 phosphate after V5+ treatment. Influence of V5+ on the levels of fructose 2, 6 phosphate, glucosamine 6 phosphate and glucose 1, 6 phosphate (HPLC), and polyphosphates, UDPG and ATP (31P NMR) was also established. Increase of sugar phosphates content was not observed after addition of vanadyl (V4+), indicating that only vanadate influences its metabolism. Obtained results from in vivo experiments indicate catalytic/inhibitory vanadate action on enzymes involved in reactions of glycolysis and glycogenesis i.e., phosphoglucomutase, phosphofructokinase and glycogen phosphorylase in filamentous fungi.  相似文献   

19.
The occurrence of linear condensed polyphosphates and cyclic condensed metaphosphates was studied by means of pulse-labeling with 32P-orthophosphate (3–5 h) in a number of Phaeophyceae species: Pylaiella litoralis, Ilea fascia, Ectocarpus siliculosus and also Rhodophyceae species: Ceramium deslongchampsii, C. rubrum, Rhodomela confervoides, Porphyridium purpureum and P. aerugineum. Twodimensional cellulose thin layer chromatography revealed that in all species studied 32P-radioactivity was generally present in all oligopolyphosphates containing 2 to 7 phosphate residues, in cyclic metaphosphates (tri-, tetra-, penta- and hexametaphosphates) and in high-molecular-weight condensed phosphates which remained at the starting point. Among the low-molecular-weight condensed inorganic phosphates the trimetaphosphate had a significantly higher specific activity than the other oligophosphates which were separated on the chromatography plates as measured by the direct scanning with a Geiger-Muller counter.The phosphate uptake strongly depends on the internal pool of reserve phosphates of the algae cells. The 32P-orthophosphate incorporation of the cells is low and sluggish when growning in a synthetic medium or in sea water. Accordingly 32P appeared preferentially in the low-molecular-weight fractions of condensed phosphates since the storage phosphates were not yet used. After previous incubation in a P-free culture medium of the algae the 32P was rather rapidly incorporated and was found mostly in the highmolecular-weight condensed phosphates.During MAK-chromatography the high-molecular-weight fractions were eluted together with the nucleic acids (tRNA and DNA) while most of the low-molecular-weight fractions left the column immediately on elution.  相似文献   

20.
We have used a combination of densimetric, calorimetric, and uv absorption techniques to obtain a complete thermodynamic characterization for the formation of nucleic acid homoduplexes of known sequence and conformation. The volume change ΔV accompanying the formation of four duplexes was interpreted to reflect changes in hydration based on the electrostriction phenomenon. In 10 mM sodium phosphate buffer at pH 7, the magnitude of the measured ΔV's ranged from ?2.0 to +7.2 ml/mol base pair and followed the order of poly(rA) · poly(dT) ~ poly(dA) · poly(dT) < poly(rA) · poly(dU) ~ poly(rA) · poly(rU). Inclusion of 100 mM NaCl in the same buffer gave the range of ?17.4 to ?2.3 mL/mol base pair and the following order: poly(dA) · poly(dT) < poly(rA) · poly(dT) < poly(rA) · poly(rU) ~ poly(rA) ~ polyr(dU). Standard thermodynamic profiles of forming these duplexes from their corresponding complementary single strands indicated similar free energies that resulted from the compensation of favorable enthalpies with unfavorable entropies along with a similar counterion uptake at both ionic strengths. The differences in these compensating effects of entropy and enthalpy correlated very well with the volume change measurements in a manner suggesting that the homoduplexes in the B conformation are more hydrated than are those in the A conformation. Moreover, the increased thermal stability of these homoduplexes resulted from an increase in the salt concentration corresponding to larger hydration levels as reflected by the ΔV results. © 1993 John Wiley & Sons, Inc.  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