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 31 毫秒
1.
The interaction of poly-N6-methyladenylic acid (poly(m6A) with poly-5-bromouridylic acid (poly(BU) was studied by the mixing curve method. A.1 m6A: 2 BU stoichiometry was clearly indicated over a wide range of ionic strengths at neutral pH, while the binding of poly(m6A) to poly(U) is known to occur with 1 m6A:1 U. Digestion by nuclease S1 confirmed this stoichiometry, indicating the absence of single strands in a 1:2 mixture. Heating profile analysis and hydroxyapatite column chromatography provided further confirmation of this finding. To determine whether 1:2 stoichiometry holds in a monomer-polymer system, the interaction of N6-methyl-9-methyladenine (m6m9A), a corresponding monomer of poly(m6A), with poly(BU) was investigated. Equilibrium dialysis experiments showed the stoichiometry of the interaction to be 1 m6A:2 BU. Thus, we would describe some structural studies of the above complexes using c.d. and i.r. spectroscopy. Poly (m6A).2poly(BU) and m6m9A.2poly(BU) are helical and analogous to each other in structure, and the bases in the complexes are all bound by hydrogen-bonding. N6-(delta 2-isopentenyl)- and N6-allyl-9-methyladenine were also found to form complexes with poly(BU), giving similar c.d. spectra with that of m6m9A.2poly(BU). The melting experiments indicated the Tms to be substantially decreased, compared to the parent unmodified complexes, even though the Tm dependence of the polymer complex on salt concentration conforms to the typical triple strand. In the following, the biological significance of this novel pairing will be discussed.  相似文献   

2.
Complexes of poly(BU) with various adenine derivatives were investigated by circular dichroism (CD) and absorption spectroscopy. A 1:2 stoichiometry was indicated on CD mixing curves for typical complexes of 9-substituted adenine and 2-aminoadenine derivatives with poly(BU). The CD spectrum of adenosine·2poly(BU) is characterized by well-resolved bands in the range of 210–350 nm. Other adenine derivative–poly(BU) complexes also afford similar CD spectra, while 2-aminoadenine derivative–poly(BU) complexes give quite different spectra. Attempts to assign representative CD spectra were made using the transition of helical poly(BU) and the respective purine polynucleotides. The similarity of the CD spectra suggests that poly(A)·2poly(BU) and adenine derivative–poly(BU) complexes are nearly identical in structure except for the ribose–phosphate linkage. The fact that the uv isosbestic point of adenosine·2poly(BU) falls in close proximity to that of the corresponding polymer complex also supports this conclusion. In the formation of stable helices, the ribose moiety is dispensable in the “strand” of purine. The Tm of 9-methyladenine·2poly(BU) is somewhat higher than that of adenosine·2poly(BU) under equivalent conditions. The Tm difference with the monomer–poly(U) system was found to be about 20°C in 0.4M NaCl–0.02M Na–cacodylate–5 × 10?4M EDTA (pH 7.0). Further, it was noted that the monomer–poly(BU) complexes are formed even when the Tm is lower than that of self-folded poly(BU).  相似文献   

3.
The interaction of poly-5-bromouridylic acid [poly(BU)] with adenosine and 9-methyladenine was studied by equilibrium dialysis, optical melting, and microcalorimetry. The stacking free energy, ω, was estimated as ?17.6 kJ/mol for adenosine·2poly(BU) and ?18.8 kJ/mol for 9-methyladenine·2poly(BU) from the binding isotherms constructed from equilibrium dialysis results. The binding isotherms constructed from a series of melting curves also gave ω values for adenosine·2poly(BU). The thermal stability of the complex depends on monomer concentration, and the partial molar enthalpies of the complex formation at the midpoint of the transition were evaluated from the Tm coefficients as a function of free monomer concentration. The values of ?92.0 and ?90.4 kJ/mol were obtained for adenosine·2poly(BU) and 9-methyladenine·2poly(BU) in 0.4M NaCl–0.02M Na-cacodylate–5 × 10?4M EDTA (pH 7.0), respectively. Microcalorimetric measurements provided lower integral heats of reaction values for these complexes, i.e., ?73.2 kJ/mol for adenosine·2poly(BU) and ?71.5 kJ/mol for 9-methyladenine·2poly(BU). A comparison with a polyribouridylic acid system provided a quantitative understanding of a stabilization by bromination in terms of thermodynamic parameters.  相似文献   

4.
Thermodynamics of base interaction in (A)n and (A.U)n   总被引:2,自引:0,他引:2  
Using precision scanning microcalorimetry we studied (A)n and (A·U)n melting in highly diluted solutions (0.3 to 5.0 mm) with different Na+ activity. This permitted us to determine directly the thermodynamic functions of stacking interaction in (A)n and base-pairing in (A·U)n. For (A-A) stacking at (A)n melting temperature we obtained ΔH(A)nm = 12.6 kJ mol?1; ΔS(A)nm = 41 J K?1 mol?1. For A·U base-pairing at a standard temperature of 298 K and 0.1 m-Na+ we have: ΔH(A·U) = 34 kJ mol?1; ΔS(A·U) = 102 J K?1 mol?1ΔG(A·U) = ?3.5 kJ mol?1.  相似文献   

5.
Raman studies of nucleic acids. VII. Poly A-poly U and poly G-poly C   总被引:6,自引:0,他引:6  
L Lafleur  J Rice  G J Thomas 《Biopolymers》1972,11(12):2423-2437
Laser-excited Raman spectra of the double-helical complexes poly A·poly U and poly G·poly C are reported for 2H2O and H2O solutions. The spectra are discussed in relation to their use as quantitative reference spectra for determining the dependence of the Raman scattering of RNA on secondary structure. The Raman line at 815 cm?1, due to the phosphodiester group, exhibits the same intrinsic intensity in spectra of poly A·poly U and poly G·poly C and is thus dependent only upon the amount of ordering of the helix and not on the kinds of nucleotides involved. The hypochromic Raman lines in spectra of poly A·poly U are identified and their intensity changes are determined quantitatively over the temperature range 32–85°C. Comparison of the spectra in the 1500–1750 cm?1 region reveals that the Raman lines from carbonyl group vibrations of uracil are about sevenfold more intense than those of guanine and cytosine for both paired and unpaired states and will thus dominate the spectra of RNA. The Raman frequencies in this region are also compared with previously reported infrared frequencies and give evidence of being strongly perturbed by base-stacking interactions in the helices.  相似文献   

6.
The N6-methyladenine and 5-methylcytosine contents in the DNA of bacteriophages M13 and fd have been analyzed. The results are summarized as follows. (1) After growth in bacteria harboring the N-3ft? drug resistance-factor, fd and M13 are observed to contain approximately 1 to 2 more 5-methylcytosine residues per DNA molecule than after growth in the parental drug-sensitive host; no effect on the N6-methyladenine content is produced by the plasmid. (2) After growth in bacteria harboring P1 prophage, fd and M13 are observed to contain approximately 2 to 3 more N6-methyladenine residues per DNA molecule than after growth in the parental P1-sensitive host; no apparent effect on the 5-methylcytosine content was produced by the P1 plasmid. (3) In agreement with others, fd carrying B-host specificity (fd·B) is observed to contain 2 more N6-methyladenine residues/DNA molecule than fd·K.  相似文献   

7.
Melting parameters of 2U:1A complexes formed by polyuridylic acid [poly(U)] and three adenine dinucleotides, diribonucleoside monophosphonate ApA and diastereoisomers of dideoxyribonucleoside methyl phosphonate [(dApA)1 and (dApA)2], in 1M NaCl and at a number of dinucleotide concentrations were obtained from differential scanning microcalorimetric data and interpreted in terms of the theory of helix–coil equilibrium in oligonucleotide–polynucleotide systems. The apparent binding constant, 1/cm, at 39°C and melting temperatures, Tm, at 1 × 10?3 M dinucleotide concentration indicate the following order of thermodynamic stability of the complexes: 2 poly(U) · (dApA)2 (2.27 × 103M?1, 44.2°C) > 2 poly(U) · (dApA)1 (9.9 × 102M1, 39.2°C) > 2 poly(U) · (ApA) (5.9 × 102M?1, 35.8°C). Corresponding calorimetric enthalpies of melting, ΔHm: 13.5, 12.7, and 12.8 kcal/mol (UUA base triplets) were found to be considerably lower than the van't Hoff enthalpies, ΔHapp: 29.4, 16.2, and 16.2 kcal/mol, respectively, evaluated from the dependence of the melting temperatures on dinucleotide concentration. Self-association of dinucleotides and their simultaneous binding as monomers, dimers, and higher-order associated species is suggested as the most probable cause of the differences between ΔHm and ΔHapp values. The differences in thermodynamic properties of the complexes formed by (dApA)1 and (dApA)2 diastereoisomers are discussed in connection with their known conformational properties. The higher and essentially enthalpic stability of the 2 poly(U) · (dApA)2 complex correlates with a lower degree of intramolecular stacking of the (dApA)2 isomer. The hydrophobically enhanced strong self-association of the latter greatly influences the thermodynamics of its complex formation with poly(U) and results in ΔHappHm = 2.3.  相似文献   

8.
 As an extension of our earlier discoveries that ZnII-cyclen complex (1) (cyclen=1,4,7,10-tetraazacyclododecane) and ZnII-acridine-pendant cyclen complex ZnII-N-(9-acridin)ylmethyl-cyclen (3) are the first compounds to selectively recognize thymidine and uridine nucleosides in aqueous solution at physiological pH, the interaction of these and a relevant complex, bis(ZnII-cyclen) (7), has been investigated with a series of polynucleotides, single-stranded poly(U) and poly(G), and double-stranded poly(A)·poly(U), poly(dA)·poly(dT) and poly(dG)·poly(dC). These ZnII-cyclen complexes interact with the imide-containing nucleobases in the single-stranded poly(U), unperturbed by the presence of the anionic phosphodiester backbone. The affinity constant of 1 for each N(3)-deprotonated uracil base in poly(U) is determined to be log K= 5.1 by a kinetic measurement, which is almost the same as log K=5.2 for the interaction of 1 with uridine. Thus, they disrupt the A-U (or A-T) hydrogen bonds to unzip the duplex of poly(A)·poly(U) or poly(dA)·poly(dT), as demonstrated by lowering of the melting temperatures (T m) of poly(A)·poly(U) and poly(dA)·poly(dT) in 5 mM Tris-HCl buffer (pH 7.6, 10 mM NaCl) with increase in their concentrations. The order of the denaturing efficiency is well correlated with that of the 1 : 1 affinity constants for each complex with uracil or thymine;7>3>1. The comparison of circular dichroism (CD) spectra for poly(A)·poly(U), poly(A), and poly(U) in the presence of 3 has revealed a structural change from poly(A)·poly(U) to two single strands, poly(A) and poly(U), caused by 3 binding exclusively to uracils in poly(U). On the other hand, the acridine-pendant cyclen complex 3, which earlier was found to associate with guanine by the ZnII coordinating with guanine N(7), in addition to the π-π stacking, interacts with guanine in the double helix of poly(dG)·poly(dC) from outside and stabilized the double-stranded structure, as indicated by higher T m. Received: 31 December 1997 / Accepted: 23 February 1998  相似文献   

9.
Temperature-dependent conformational transitions of spin-labeled poly rA, spin-labeled poly rU and the two-stranded helical complexes consisting either of spin-labeled rA·poly rU or spin-labeled poly rU·poly rA have been measured by electron spin resonance spectrocopy. The polynucleotides were spin labeled with 4-(2-iodoacetamido)2,2,6,6-tetramethylpiperidinooxyl and the spin label to nucleotide base ratio was approximately 1:600. The relationship between the log of tumbling time τ and the reciprocal absolute temperature for the spin-labeled single and double-stranded polynucleotides is presented. An agreement between TmOD (optical density melting) and Tmsp (spin melting) is found for the complexes, which strongly supports the conclusion that the same temperature-dependent structural changes are monitored with both techniques.  相似文献   

10.
Experiments were designed to test for functional differences which might shed light on the differences in actin-activated ATPase activities recently reported for myosin subfragments-1 bearing different light chains. By using the method of A. G. Weeds and R. S. Taylor (1975, Nature (London)257, 54), two types of subfragment-1 (S-1) from myosin of rabbit fast skeletal muscle were prepared: (S-1)·A1 and (S-1)·A2 bearing, respectively, the alkali-1 and alkali-2 light chains. (In agreement with the findings of these investigators, actin enhanced the ATPase activity of (S-1)·A1 more than that of (S-1)·A2 at lower actin concentrations.) Through use of time-resolved fluorescence depolarization techniques, the affinity constants for the binding of the two types of S-1 to F-actin in the absence of ATP were found to be very similar: 3.4 ± 0.3 × 106m?1 (N = 10) for (S-1)·A1 and 3.9 ± 0.2 × 106m?1 (N = 7) for (S-1)·A2 of one preparation, and 6.4 ± 0.2 × 106m?1 (N = 6) for (S-1)·A1 and 7.7 ± 0.5 × 106m?1 (N = 12) for (S-1)·A2 of another preparation (pH 7.0, 25 °C, 0.28 m KCl, 1.5 mm MgCl2, 0.5 mm ethylene glycol bis (β-aminoethyl ether) N,N′-tetracetic acid, 10 mm imidazole, and 0.1 mmN-tris (hydroxymethyl) methyl-2-aminoethane sulfonate). The affinity constants for the two species of S-1 and actin also have a similar dependence on ionic strength and are not affected by addition of 0.6 mm CaCl2 to the above solution. The CaATPase (or the CaITPase) activities of the two species of S-1 show the same pH dependence.  相似文献   

11.
Isotherms of the EtBr adsorption on native and denatured poly(dA)poly(dT) in the temperature interval 20–70°C were obtained. The EtBr binding constants and the number of binding sites were determined. The thermodynamic parameters of the EtBr intercalation complex upon changes of solution temperature 20–48°C were calculated: 1.0·106 M−1K≤1.4·106 M−1, free energy ΔG o=−8.7±0.3 kcal/mol, enthalpy ΔH o≅0, and entropy ΔS o=28±0.5 cal/(mol deg). UV melting has shown that the melting temperature (T m) of EtBr-poly(dA)poly(dT) complexes (μ=0.022,4.16·10−5 M EtBr) increased by 17°C as compared with the ΔT m of free homopolymer, whereas the half-width of the transition (T m) is not changed. It was shown for the first time that EtBr forms complexes of two types on single-stranded regions of poly(dA)poly(dT) denatured at 70°C: strong (K 1=1.7·105 M−1; ΔG o=−8.10±0.03 kcal/mol) and weak (K 2=2.9·103 M−1; ΔG o=−6.0±0.3 kcal/mol).The ΔG o of the strong and weak complexes was independent of the solution ionic strength, 0.0022≤μ≤0.022. A model of EtBr binding with single-stranded regions of poly(dA)poly(dT) is discussed.  相似文献   

12.
The synthesis of complexes of Li(I), K(I), Mg(II), Ca(II) and Ba(II) with guanosine in basic non aqueous solutions is described. The complexes were of two types: (1) complexes having the general formula, M(Guo)nXm·YH2O·ZC2H5OH, where M = Mg(II), Ca(II), Ba(II) and Li(I), n = 1,2,4, X = Cl?, Br?, NO3?, ClO4? and OH?, m = 1,2, Y = 0?6 and X = 0?2, and (2) complexes with the general formula, M(GuoH-1)(OH)n?1·YH2O, where M = K(I), Ca(Il) and Ba(II), GuoH-1 =Ionized guanosine at N1, n = 1,2 and Y = 1?3. The complexes are characterized by their proton nuclear magnetic resonance (1H NMR) and Fourier transform infrared (FT-IR) spectra. The FT-IR and 1H NMR data of the non ionized nucleoside complexes suggest that the metal binding is through the N7-site of guanine and that the anion (X) is hydrogen bonded to N1H and NH2 groups. In the N1-ionized guanosine complexes the metal binding is via the O6? of guanine. All the complexes formed exhibited a transition of the sugar conformation from C2-endo/anti in the free nucleoside to C3-endo/anti in the metal complexes.  相似文献   

13.
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   

14.
The conformational properties of 5-fluorouracil derivatives are compared to uracil derivatives. FUrd, 5′-FUMP, and poly(FU) are studied as a function of pH and temperature by 19F- and 1H-nmr spectroscopy, and the corresponding uracil derivatives by 1H-nmr spectroscopy. FUrd exhibits no significant conformational changes with solution pH (5–10). In contrast, at low pH (6–7) 5′-FUMP and 5′-UMP show similar conformational features, while at high pH (9) 5′-FUMP shows significant conformational alterations. Also, poly(U) and poly(FU) are conformationally similar at low pH, but increasing pH induces changes in poly(FU). These changes are observed in the backbone [γ(C4′-C5′)], furanose, and furanose-base conformations. The apparent pKa of N3-H ionization of the FUra base is determined by 1H- and 19F-nmr to range from 7.5 to 8.2 [FUrd < 5′-FUMP < 5′-FUDP < poly(FU)]. These observations are interpreted as a result of electrostatic interactions generated between the ionized phosphate group and the negatively charged base moiety as the pH is raised. The interaction properties of poly(FU) with ApA are studied by 1H- and 19F-nmr spectroscopy, and these properties compared to those published for poly(U). Poly(FU) forms a complex with ApA inducing upfield 1H-shifts in both components, and downfield 19F- shifts in poly(FU). The base stoichiometry of the complex for poly(U)·ApA is 2U:1A at various U/A ratios. In contrast, the base stoichiometry of the poly(FU)·ApA complex appears to be dependent on the FU/A ratio. At high FU/A ratio, the complex is 2FU:1A, and as the FU/A ratio approaches unity the complex becomes 1FU:1A.  相似文献   

15.
Methyltransferases from the m1A58 tRNA methyltransferase (TrmI) family catalyze the S-adenosyl-l-methionine-dependent N1-methylation of tRNA adenosine 58. The crystal structure of Thermus thermophilus TrmI, in complex with S-adenosyl-l-homocysteine, was determined at 1.7 Å resolution. This structure is closely related to that of Mycobacterium tuberculosis TrmI, and their comparison enabled us to enlighten two grooves in the TrmI structure that are large enough and electrostatically compatible to accommodate one tRNA per face of TrmI tetramer. We have then conducted a biophysical study based on electrospray ionization mass spectrometry, site-directed mutagenesis, and molecular docking. First, we confirmed the tetrameric oligomerization state of TrmI, and we showed that this protein remains tetrameric upon tRNA binding, with formation of complexes involving one to two molecules of tRNA per TrmI tetramer. Second, three key residues for the methylation reaction were identified: the universally conserved D170 and two conserved aromatic residues Y78 and Y194. We then used molecular docking to position a N9-methyladenine in the active site of TrmI. The N9-methyladenine snugly fits into the catalytic cleft, where the side chain of D170 acts as a bidentate ligand binding the amino moiety of S-adenosyl-l-methionine and the exocyclic amino group of the adenosine. Y194 interacts with the N9-methyladenine ring, whereas Y78 can stabilize the sugar ring. From our results, we propose that the conserved residues that form the catalytic cavity (D170, Y78, and Y194) are essential for fashioning an optimized shape of the catalytic pocket.  相似文献   

16.
N6-methyladenine (m6A) is a rare base naturally occurring in DNA. It is different from the base adenine due to its N-CH3. Therefore, the base not only pairs with thymine, but also with other DNA bases (cytosine, adenine and guanine). In this work, Møller-Plesset second-order (MP2) method has been used to investigate the binding mechanism between m6A and natural DNA bases in gas phase and in aqueous solution. The results show that N-CH3 changed the way of N6-methyladenine binding to natural DNA bases. The binding style significantly influences the stability of base pairs. The trans-m6A:G and trans-m6A:C conformers are the most stable among all the base pairs. The existence of solvent can remarkably reduce the stability of the base pairs, and the DNA bases prefer pairing with trans-m6A to cis-m6A. Besides, the properties of these hydrogen bonds have been analyzed by atom in molecules (AIM) theory, natural bond orbital (NBO) analysis and Wiberg bond indexes (WBI). In addition, pairing with m6A decreases the binding energies compared to the normal Watson-Crick base pairs, it may explain the instability of the N6 site methylated DNA in theory.
Figure
Figure The most stable configurations of the base pairs  相似文献   

17.
Abstract

Parallel-stranded (ps) DNA hairpins with alternating d(A-isoG)/d(T·C) (designated as ps-t1) and d(A·G)/d(T·m5isoC) (ps-t2) sequences were studied by means of UV, CD and fluorescence spectroscopy. The thermostability of d(A·G)/d(T·m5isoC) sequence was close to that of aps d(G·A)/d(T·C). The stability of the ps d(A·isoG)/d(T·C) sequence was even higher than that of a related anti-parallel-stranded (aps) d(G·A)/d(T·C) sequence, being unique for ps DNAs studied so far.  相似文献   

18.
This study describes the relationships between dinitrogen (N2) fixation, dihydrogen (H2) production, and electron transport associated with photosynthesis and respiration in the marine cyanobacterium Trichodesmium erythraeum Ehrenb. strain IMS101. The ratio of H2 produced:N2 fixed (H2:N2) was controlled by the light intensity and by the light spectral composition and was affected by the growth irradiance level. For Trichodesmium cells grown at 50 μmol photons · m?2 · s?1, the rate of N2 fixation, as measured by acetylene reduction, saturated at light intensities of 200 μmol photons · m?2 · s?1. In contrast, net H2 production continued to increase with light levels up to 1,000 μmol photons · m?2 · s?1. The H2:N2 ratios increased monotonically with irradiance, and the variable fluorescence measured using a fast repetition rate fluorometer (FRRF) revealed that this increase was accompanied by a progressive reduction of the plastoquinone (PQ) pool. Additions of 2,5‐dibromo‐3‐methyl‐6‐isopropyl‐p‐benzoquinone (DBMIB), an inhibitor of electron transport from PQ pool to PSI, diminished both N2 fixation and net H2 production, while the H2:N2 ratio increased with increasing level of PQ pool reduction. In the presence of 3‐(3,4‐dichlorophenyl)‐1,1‐dimethylurea (DCMU), nitrogenase activity declined but could be prolonged by increasing the light intensity and by removing the oxygen supply. These results on the coupling of N2 fixation and H2 cycling in Trichodesmium indicate how light intensity and light spectral quality of the open ocean can influence the H2:N2 ratio and modulate net H2 production.  相似文献   

19.
N6‐methyladenine is the most widespread mRNA modification. A subset of human box C/D snoRNA species have target GAC sequences that lead to formation of N6‐methyladenine at a key trans Hoogsteen‐sugar A·G base pair, of which half are methylated in vivo. The GAC target is conserved only in those that are methylated. Methylation prevents binding of the 15.5‐kDa protein and the induced folding of the RNA. Thus, the assembly of the box C/D snoRNP could in principle be regulated by RNA methylation at its critical first stage. Crystallography reveals that N6‐methylation of adenine prevents the formation of trans Hoogsteen‐sugar A·G base pairs, explaining why the box C/D RNA cannot adopt its kinked conformation. More generally, our data indicate that sheared A·G base pairs (but not Watson–Crick base pairs) are more susceptible to disruption by N6mA methylation and are therefore possible regulatory sites. The human signal recognition particle RNA and many related Alu retrotransposon RNA species are also methylated at N6 of an adenine that forms a sheared base pair with guanine and mediates a key tertiary interaction.  相似文献   

20.
This study is concerned with the isolation and characterization of the enzyme, S-adenosylmethionine:ribosomal ribonucleic acid-adenine (N6−) methyl-transferase [rRNA-adenine (N6-) methylase] of Escherichia coli strain B, which is responsible for the formation of N6-methyladenine moieties in ribosomal ribonucleic acids (rRNA). A 1,500-fold purified preparation of the species-specific methyltransferase methylates a limited number of adenine moieties in heterologous rRNA (Micrococcus lysodeikticus and Bacillus subtilis) and methyl-deficient homologous rRNA. The site recognition mechanism does not require intact 16 or 23S rRNA. The enzyme does not utilize transfer ribonucleic acid as a methyl acceptor nor does it synthesize 2-methyladenine or N6-dimethyladenine moieties. Mg2+, spermine, K+, and Na+ increase the reaction rate but not the extent of methylation; elevated concentrations of the cations inhibit markedly. The purified preparations utilize 9-β-ribosyl-2,6-diaminopurine (DAPR) as a methyl acceptor with the synthesis of 9-β-ribosyl-6-amino-2-methylaminopurine. A comparison of the two activities demonstrated that one methyltransferase is responsible for the methylation of both DAPR and rRNA. This property provides a sensitive assay procedure unaffected by ribonucleases and independent of any specificity exhibited by rRNA methyl acceptors.  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