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 contemporary genetic code and the process of protein biosynthesis most assuredly evolved from a simpler code and process. We believe that there was obligatory coevolution of the two and that the earlier code and process must have involved a more direct linkage between the amino acids and the informational macromolecule. We propose that an early form of translating existed in which amino acids were attached directly to the ‘messenger’ RNA along the backbone as 2'OH aminoacyl esters. These esters then condensed with each other on the RNA backbone yielding a peptide covalently attached to the RNA, without the use of tRNAs and ribosomes. This presentation is concerned with experimental data which indicate that such a simple translation system is possible and must have involved the following steps: (1) formation of the aminoacyl adenylate anhydride, (2) transfer of the amino acid from the adenylate to imidazole, (3) transfer of the amino acid from imidazole to 2'OH groups along the backbone of RNAs (4) condensation of the amino acids to yield peptides. Steps (1)–(3) have been confirmed in chemical systems. Our preliminary evidence indicates step (4) is also possible. The aminoacylation of polyribonucleotides and the subsequent formation of peptides is a dynamic and experimentally accessible system for studying genetic coding specifities and our present studies are now concentrated on step (4), looking for such specifities.  相似文献   

2.
One-dimensional nuclear Overhauser effect (NOE) in nuclear magnetic resonance spectroscopy along with stereochemically sound model building was employed to derive the structure of the hybrid poly(rA).poly(dT) in solution. Extremely strong NOE was observed at AH2' when AH8 was presaturated; strong NOEs were observed at TH2'TH2' when TH6 was presaturated; in addition the observed NOEs at TH2' and TH2' were nearly equal when TH6 was presaturated. There was no NOE transfer to AH3' from AH8 ruling out the possibility of (C-3'-endo, low anti chi approximately equal to 200 degrees to 220 degrees) conformation for the A residues. The observed NOE data suggest that the nucleotidyl units in both rA and dT strands have equivalent conformations: C-2'-endo/C-1'-exo, anti chi approximately equal to 240 degrees to 260 degrees. Such a nucleotide geometry for rA/dT is consistent with a right-handed B-DNA model for poly(rA).poly(dT) in solution in which the rA and dT strands are conformationally equivalent. Molecular models were generated for poly(rA).poly(dT) in the B-form based upon the geometrical constraints as obtained from the NOE data. Incorporation of (C-2'-endo pucker, chi congruent to 240 degrees to 260 degrees) into the classical B-form resulted in severe close contacts in the rA chain. By introducing base-displacement, tilt and twist along with concomitant changes in the backbone torsion angles, we were able to generate a B-form for the hybrid poly(rA).poly(dT) fully consistent with the observed NOE data. In the derived model the sugar pucker is C-1'-exo, a minor variant of C-2'-endo and the sugar base torsion is 243 degrees, the remaining torsion angles being: epsilon = 198 degrees, xi = 260 degrees, alpha = 286 degrees, beta = 161 degrees and gamma = 72 degrees; this structure is free of any steric compression and indicates that it is not necessary to switch to C-3'-endo pucker for rA residues in order to accommodate the 2'-OH group. The structure that we have proposed for the polynucleotide RNA-DNA hybrid in solution is in complete agreement with that proposed for a hexamer hybrid in solution from NOE data and is inconsistent with the heteronomous model proposed for the fibrous state.  相似文献   

3.
Left-handed Helical Polynucleotides with D-Sugar Phosphodiester Backbones   总被引:3,自引:0,他引:3  
Naturally occurring polynucleotides have right-handed helical confrontations in the solid state1 and in solution2. Poly(dI-dC)poly(dI-dC) was found to form a left-handed helix in spite of the D-sugar backbone. Also, L-adenylyl-(3′–5′)-L-adenosine synthesized by Tazawa et al4. takes up the left-handed stacked conformation. We had synthesized a dinucleoside monophosphate, 8,2′-anhydro-8-mercapto-9-β-D-arabinofuranosyladenine phosphoryl-(3′–5′)-8,2′-anhydro-8-mercapto-9-β-D-arabinofuranosyladenine (AspAs) (molecular structure Ia; see also ref. 5) and this compound has a left-handed stacked conformation. The two bases in Ia, having the D-sugar backbone, stacked along the left-handed helical axis; these bases are fixed at ?CN = ?108° (syn-anti region) by the anhydro linkages.  相似文献   

4.
B. Pispisa  M. Farinella 《Biopolymers》1984,23(8):1465-1479
The oxidation of L -adrenaline (epinephrine) in the presence of [Fe(tetpy)(OH)2]+ ions bound to poly(L -glutamate) or to poly(D-glutamate) has been studied at pH 7 (tetpy = 2,2′:6′,2″:6″,2?-tetrapyridyl). Electron transfer from the substrate to the central metal ion, which is rate-determining, proceeds stereoselectively only when extensive and possibly specific interactions between adrenaline and the peptidic residues of the ordered polymer in the close environment of the active sites occur. This ensures different steric constraints for the two diastereomeric precursor complexes, which are thought to affect the separation and orientation of the redox centers differently, leading to the observed phenomena. Some data on the catalytic oxidation of L -dopa(3,4-dihydroxyphenylalanine) are also presented, showing stereoselective effects similar to those observed with L -adrenaline, despite the diverse distance of the chiral center from the reacting OH groups. A mechanistic interpretation of the results is discussed in the light of a few general considerations concerning the structural features of the catalytic systems. Possible explanations for the finding that stereoselectivity occurs at the expense of the efficiency of catalysis are also considered.  相似文献   

5.
Abstract

ID NOE 1H NMR spectroscopy at 500 MHz was employed to examine the structure of poly(dA)·poly(dT) in solution. NOE experiments were conducted as a function of presaturation pulse length (50, 30, 20 and 10 msec) and.power (19 and 20 db) to distinguish the primary NOEs from spin diffusion. The 10 msec NOE experiments took 49 hrs and over 55,000 scans for each case and the difference spectra were almost free from diffusion.

The spin diffused NOE difference spectra as well as difference NOE spectra in 90% H2O + 10% D2O in which TNH3 was presaturated enabled to make a complete assignment of the base and sugar protons. It is shown that poly(dA) ·poly(dT) melts in a fashion in which single stranded bubbles are formed with increasing temperature.

Extremely strong primary NOEs were observed at H2′/H2″ when AH8 and TH6 were presaturated. The observed NOEs at AH2′ and that AH2″ were very similar as were the NOEs at TH2′ and TH2″. The observed NOEs at AH2′ and AH2″when AH8 was presaturated were very similar to those observed at TH2′ and TH2″ when TH6 was presaturated. In addition, presaturation of H1′ of A and T residues resulted in similar NOEs at AH2′/H2″ and TH2′/H2″ region and these NOEs at H2′ and H2″ were distinctly asymmetric as expected in a C2′-endo sugar pucker. There was not a trace of NOE at AH8 and TH6 when AH3′ and TH3′ were presaturated indicating that C3′-endo, × = 30–40° conformation is not valid for this DNA. From these NOE data, chemical shift shielding calculations and stereochemistry based computer modellings, we conclude that poly(dA)·poly(dT) in solution adopts a right- handed B-DNA duplex in which both dA and dT strands are conformationally equivalent with C2′-endo sugar pucker and a glycosyl torsion, ×, of ?73°, the remaining backbone torsion angles being φ′ = 221°, ω′ = 212°, ω = 310°, φ = 149°, ψ = 42°, ψ′ = 139°. The experimental data are in total disagreement with the heteronomous DNA model of Arnott et. al. proposed for the fibrous state. (Arnott, S., Chandrasekaran, R., Hall, I.H., and Puigjaner, L.C., Nucl. Acid Res. 11, 4141, 1983).  相似文献   

6.
The helix–coil transition and conformational structure of poly(8-bromoadenylic acid) [poly(8BrA)] have been investigated using 1H- and 13C-nmr, CD, and ir spectroscopy. The results have been compared with the structure of the related 5′-mono- and polynucleotides. The chemical shifts of H(2′), H(3′), C(2′), and C(3′) nmr signals show an interesting correlation with both the puckering of ribose ring and glycosidic bond torsion angle. Poly(8BrA) shows an upfield shift of the C(3′) signal and a downfield shift of the H(3′) signal compared to the chemical shifts in poly(A). These shifts are consistent with a C(3′) endo-syn conformation for poly(8BrA). A similar effect has been reported previously and is also observed here on the C(2′) and H(2′) signals when the preferred conformation is C(2′)endo-syn (e.g., in 5′-8BrAMP). The chemical-shift parameters thus act as a probe for studying syn ? anti and N ? S equilibria in solutions. The three-bond 1H-′13C coupling constants between H(1′) and C(8) and C(4) have been measured in poly(8BrA) and 5′-8BrAMP and their structural implications have been discussed. The observed preference of a C(3′)endo-syn conformation for poly(8BrA), coupled with other evidence, throws doubt on the validity of a correlation previously reported whereby a syn conformation is associated with a C(2′)endo ribose pucker. The backbone conformation of randomly coiled poly(8BrA) is very similar to the structures found in polyribonucleotides: poly(A) and poly(U). All three polymers show strong preferences for the backbone angles found in RNA helices. The CD spectrum of poly(8BrA) has a striking relationship to that of poly(A). The signs of all extrema are inverted, and the magnitudes are related by a constant factor. We suggest that these differences result from a change in the angle between coupled transition moment vectors in the two polymers. Infrared spectra of poly(8BrA) in H2O and D2O solution are reported for the frequency range below 1400 cm?1. The antisymmetric >PO stretching vibration is observed at an unusually low frequency in the helix (1214 cm?1). The symmetric >PO stretch occurs at ~1095 cm?1 but is not resolved from a ring vibration near this frequency. A conformationally sensitive band, characteristic of helical RNA structures, is observed at 817 cm?1 and disappears when the helix is melted. This observation confirms the conclusion that ordered poly(8BrA) has a regular helical structure with an RNA backbone conformation. A stereochemical explanation is provided for the failure of poly(8BrA) (or other syn polymers) to form double helices with anti-polyribonucleotides.  相似文献   

7.
We report on an unrestrained molecular dynamics simulation of the flavin mononucleotide (FMN)–RNA aptamer. The simulated average structure maintains both cross‐strand and intermolecular FMN–RNA nuclear Overhauser effects from the nmr experiments and has all qualitative features of the nmr structure including the G10–U12–A25 base triple and the A13–G24, A8–G28, and G9–G27 mismatches. However, the relative orientation of the hairpin loop to the remaining part of the molecule differs from the nmr structure. The simulation predicts that the flexible phosphoglycerol part of FMN moves toward G27 and forms hydrogen bonds. There are structurally long‐lived water molecules in the FMN binding pocket forming hydrogen bonds within FMN and between FMN and RNA. In addition, long‐lived water is found bridging primarily RNA backbone atoms. A general feature of the environment of long‐lived “structural” water is at least two and in most cases three or four potential acceptor atoms. The 2′‐OH group of RNA usually acts as an acceptor in interactions with the solvent. There are almost no intrastrand O2′H(n)⋮O4′(n + 1) hydrogen bonds within the RNA backbone. In the standard case the preferred orientation of the 2′‐OH hydrogen atoms is approximately toward O3′ of the same nucleotide. However, a relatively large number of conformations with the backbone torsional angle γ in the trans orientation is found. A survey of all experimental RNA x‐ray structures shows that this backbone conformation occurs but is less frequent than found in the simulation. Experimental nmr RNA aptamer structures have a higher fraction of this conformation as compared to the x‐ray structures. The backbone conformation of nucleotide n + 1 with the torsional angle γ in the trans orientation leads to a relatively short distance between 2′‐OH(n) and O5′(n + 1), enabling hydrogen‐bond formation. In this case the preferred orientation of the 2′‐OH hydrogen atom is approximately toward O5′(n + 1). We find two relatively short and dynamically stable types of backbone–backbone next‐neighbor contacts, namely C2′(H)(n)⋮O4′(n + 1) and C5′(H)(n + 1)⋮O2′(n). These interactions may affect both backbone rigidity and thermodynamic stability of RNA helical structures. © 1999 John Wiley & Sons, Inc. Biopoly 50: 287–302, 1999  相似文献   

8.
In the IR spectra, the coupling of vibrations leads to band splitting and/or bands shifting in opposite directions which provides information on the mutual orientation of groupings. From such band shifts in the range 1800 to 1500 cm-1 one can draw conclusions on the double helix formation of polynucleotides. These band shifts are caused either by vibrational coupling of stretching vibrations within pairs of base residues or by coupling of stretching vibrations with the bending (scissor) vibration of the -NH2 groups; the latter is indicated by band shifts after deuterium substitution within the amino groups. Couplings of phosphate and 1 ibose vibrations in the range 1300 to 1000 cm-1 provide information on the secondary structure of the backbone. In order to obtain information of the structure of the RNA backbone, the IR spectra of poly(ribonucleotides) were studied in neutral media in which they were single-stranded. The shift due to coupling of the band of the 2'OD bending vibration and that of the antisymmetric stretching vibration of the ether group of the ribose residue proves that ribose residues of the backbone are cross-linked via hydrogen bonds. These are formed between the 2'OD or 2'OH groups, respectively, and the O atoms of the ether group of the neighboring ribose residues. This is the reason for the difference between DNA and RNA as regards the 2'OH group. The structure formation caused by these hydrogen bonds results in a stiffening of the RNA backbone. The tendency to form these hydrogen bonds increases in the order poly (U), poly(C), poly (A). This order of secondary structure stabilization is due to an interplay between the influences of (1) the 2'OH hydrogen bonds and (2) the base residues' stacking. Furthermore, the coupling of the antisymmetric stretching vibration of the greater than PO2- groups with a vibration involving the 2'OH group can result in a doublet structure of the band at about 1240 cm-1 if cations with strong fields are present. This probably shows that these cations can turn the greater than PO2-groups-which are usually turned outward at the backbone, as shown by construction of molecular models- towards the basic residues. Thus they cause stiff monohelices which are right-handed screws.  相似文献   

9.
E K Achter  G Felsenfeld 《Biopolymers》1971,10(9):1625-1634
To elucidate the role of the bases in single-strand polynucleolide conformation, we have studied apurinic acid, a single-strand polydeoxyribonucleotide from which almost half the bases have been removed. The conformation of apurinic acid in aqueous solution near θ condition has been investigated by sedimentation velocity and sedimentation equilibrium measurements. The unperturbed coil dimensions of apurinic acid are essentially identical to those of poly rU of the same degree of polymerization. The dimensions are also similar to those of poly rA at high temperature, where the adenine residues are not stacked upon one another. We conclude that the considerable rigidity of these polynuclotides is conferred not by residual, undetected base stacking, but by restrictions in rotation about the bonds of the backbone. Furthermore, the rigidity of the ribose-phosphate backbone cannot be attributed to interactions involving the 2'—OH group.  相似文献   

10.
The role of non-classical hydrogen bonding in RNA structure has been investigated using polyuridylic acid, which has a labile ordered structure at temperatures near 0 °C, as a model system. By comparing the proton nuclear magnetic resonance spectrum of poly(U) in the transition region with that of uridine and the dimer UpU we find evidence that both the imino N(3)-H and the ribosyl 2′-OH protons are hydrogen bonded. The characteristics of the former are consistent with participation in N(3)-HOC bonding primarily between residues in the same strand. As yet we cannot unambiguously assign the acceptor for the 2′-OH in ordered poly(U): because of its apparent stability and the acceptable stereochemistry, we presently favor a bond between ribose 2′-OH and O(1′) connecting adjacent nucleotides of the same strand. This arrangement could contribute to the co-operativity of the poly(U) helix formation. The recently proposed 2′-OHO(1′) interactions in crystalline yeast transfer RNAPhe suggest similar interactions might play a role in the conformational stability of natural RNAs. A second conformational transition below the major transition in the ultraviolet can be detected in poly(U) by monitoring the H(6) proton of uracil.  相似文献   

11.
Guanine quadruplexes can form in both DNA and RNA and influence many biological processes through various protein interactions. The DEAD-box RNA helicase protein DDX21 has been shown to bind and remodel RNA quadruplexes but little is known about its specificity for different quadruplex species. Previous reports have suggested DDX21 may interact with telomeric repeat containing RNA quadruplex (TERRA), an integral component of the telomere that contributes to telomeric heterochromatin formation and telomere length regulation. Here we report that the C-terminus of DDX21 directly interacts with TERRA. We use, for the first time, 2D saturation transfer difference NMR to map the protein binding site on a ribonucleic acid species and show that the quadruplex binding domain of DDX21 interacts primarily with the phosphoribose backbone of quadruplexes. Furthermore, by mutating the 2′OH of loop nucleotides we can drastically reduce DDX21's affinity for quadruplex, indicating that the recognition of quadruplex and specificity for TERRA is mediated by interactions with the 2′OH of loop nucleotides.  相似文献   

12.
A tetrapetide containing an Aib residue, Boc-Asn-Aib-Thr-Aib-OMe, was synthesized as a peptide model for the N-glycosylation site in N-glycoproteins. Backbone conformation of the peptide and possible intramolecular interaction between the Asn and Thr side chains were elucidated by means of n.m.r. spectroscopy. Temperature dependence of NH proton chemical shift and NOE experiments showed that Boc-Asn-Aib-Thr-Aib-OMe has a tendency to form a β-turn structure with a hydrogen bond involving Thr and Aib4 NH groups. Incorporation of Aib residues in the peptide model promotes folding of the peptide backbone. With folded backbone conformation, carboxyamide protons of the Asn residue are not involved in hydrogen bond network, while the OH group of the Thr residue is a candidate for a hydrogen bond in DMSO-d6 solution.  相似文献   

13.
A tetrapetide containing an Aib residue, Boc-Asn-Aib-Thr-Aib-OMe, was synthesized as a peptide model for the N-glycosylation site in N-glycoproteins. Backbone conformation of the peptide and possible intramolecular interaction between the Asn and Thr side chains were elucidated by means of n.m.r. spectroscopy. Temperature dependence of NH proton chemical shift and NOE experiments showed that Boc-Asn-Aib-Thr-Aib-OMe has a tendency to form a β-turn structure with a hydrogen bond involving Thr and Aib4 NH groups. Incorporation of Aib residues in the peptide model promotes folding of the peptide backbone. With folded backbone conformation, carboxyamide protons of the Asn residue are not involved in hydrogen bond network, while the OH group of the Thr residue is a candidate for a hydrogen bond in DMSO-d6 solution.  相似文献   

14.
The effects of heteroatom substitution from a silicon atom to a germanium atom in donor‐acceptor type low band gap copolymers, poly[(4,4′‐bis(2‐ethylhexyl)dithieno[3,2‐b:2′,3′‐d]silole)‐2,6‐diyl‐alt‐(2,1,3‐benzothiadiazole)‐4,7‐diyl] (PSiBTBT) and poly[(4,4′‐bis(2‐ethylhexyl)dithieno[3,2‐b:2′,3′‐d]germole)‐2,6‐diyl‐alt‐(2,1,3‐benzothiadiazole)‐4,7‐diyl] (PGeBTBT), are studied. The optoelectronic and charge transport properties of these polymers are investigated with a particular focus on their use for organic photovoltaic (OPV) devices in blends with phenyl‐C70‐butyric acid methyl ester (PC70BM). It is found that the longer C‐Ge bond length, in comparison to C‐Si, modifies the molecular conformation and leads to a more planar chain conformation in PGeBTBT than PSiBTBT. This increase in molecular planarity leads to enhanced crystallinity and an increased preference for a face‐on backbone orientation, thus leading to higher charge carrier mobility in the diode configuration. These results provide important insight into the impact of the heavy atom substitution on the molecular packing and device performance of polymers based on the poly[2,6‐(4,4‐bis‐(2‐ethylhexyl)‐4H‐cyclopenta[2,1‐b;3,4‐b]‐dithiophene)‐alt‐4,7‐(2,1,3‐benzothiadiazole) (PCPDTBT) backbone.  相似文献   

15.
Moderately high molecular weight polycations stimulate arachidonic acid release with concomitant synthesis and release of prostaglandins in cultured 3T3 mouse fibroblasts. We have examined a series of synthetic polycations for prostaglandin synthesis-inducing activity as an approach to defining the structural features required for activity. Extensive (>80%) acetylation of poly(vinylamine) was tolerated without loss of activity, indicating that a uniform high density of charges is not required. However, complete acetylation of poly(vinylamine) abolished activity, indicating that some positive charges are required for activity. Full activity was observed for charge densities in the range of one per two to one per six atoms of polymer backbone. Branched and linear polycations activated equally well. Location of the charge with respect to the polymer backbone did not affect activity in polymers bearing charges located up to seven atoms away from the backbone. Polycations lacking primary or secondary amino groups exhibited full activity, indicating that Schiff base formation is not required for activity. These results are consistent with a model in which activation involves electrostatic interactions with discrete anionic sites on the target cell.  相似文献   

16.
Möbitz H  Boll M 《Biochemistry》2002,41(6):1752-1758
Benzoyl-CoA reductase from the anaerobic bacterium Thauera aromatica catalyzes the ATP-driven two-electron reduction of the aromatic moiety of benzoyl-CoA. A Birch mechanism involving alternate one-electron and one-proton transfer steps to the aromatic ring was previously proposed for benzoyl-CoA reductase. Due to the high redox barrier, the first electron transfer step yielding a radical anion is considered the rate-limiting step in this reaction. Focusing on the mechanism of substrate reduction, this work combines the kinetic analysis of a number of substrate analogues with a model based on the ab initio calculated electron density of the radical anion of benzoyl-CoA, a transition state model of the proposed Birch mechanism. Both K(m) and k(cat) of ortho-substituted benzoyl-CoA increased in parallel with the substituent's acceptor strength (F > Cl = H > OH > NH(2)). Among the isomers of monofluorobenzoyl-CoA, reduction rates decreased in the following order: ortho > meta > para; the K(m) values increased in the following order: meta > ortho > para. Five-ring heteroaromatic acid thiol esters were reduced in the following order: thiophene > furan > pyrrole; the 2-isomers are reduced much faster than the 3-isomers. Most of these results could be rationalized by the model. A Hammett plot indicated that the reaction mechanism is only slightly polar, suggesting the involvement of a partial protonation of the carbonyl oxygen of benzoyl-CoA and/or a simultaneous transfer of the first electron and proton. Surprisingly, benzoyl-CoA reductase exhibited a hydrogen kinetic isotope effect on k(cat) for pyridine-2-carbonyl-CoA (2.1) but only a negligible one for benzoyl-CoA (1.2), indicating that pyridine-2-carbonyl-CoA reduction proceeds according to a varied mechanism.  相似文献   

17.
The cupric complexes of poly(Nε-acetoacetyl-L -lysine), [Lys(Acac)]n′ poly(Nδ-acetoacetyl-L -ornithine), [Orn(Acac)]n′ and poly(Nγ-acetoacetyl-L -diaminobutyric acid), [A2bu-(Acac)]n, as well as of the model compound n-hexyl acetoacetamide, have been investigated by means of absorption, potentiometric, equilibrium dialysis, and CD measurements. While in the complex of the model compound, one chelating group is bound to one cupric ion, in the polymeric complexes two β-ketoamide groups are bound to Cu(II) under the same experimental conditions. The binding constant of cupric ions to the three polymers and the formation constant of the Cu(II)-nhexylacetoacetamide complex have been evluated. Investigation on the chiroptical properties of the three polymeric complexes shows that the peptide backbone does not undergo conformational transitions, remaining α-helical when up to 20% of the side chains are bound to Cu(II). The optical activity of the β-ketoamide chromophores is substantially affected by complex formation and is discussed in terms of asymmetric induction from the chiral backbone.  相似文献   

18.
Myosins in muscle assemble into filaments by interactions between the C-terminal light meromyosin (LMM) subdomains of the coiled-coil rod domain. The two head domains are connected to LMM by the subfragment-2 (S2) subdomain of the rod. Our mixed kinetic model predicts that the flexibility and length of S2 that can be pulled away from the filament affects the maximum distance working heads can move a filament unimpeded by actin-attached heads. It also suggests that it should be possible to observe a head remain stationary relative to the filament backbone while bound to actin (dwell), followed immediately by a measurable jump upon detachment to regain the backbone trajectory. We tested these predictions by observing filaments moving along actin at varying ATP using TIRF microscopy. We simultaneously tracked two different color quantum dots (QDs), one attached to a regulatory light chain on the lever arm and the other attached to an LMM in the filament backbone. We identified events (dwells followed by jumps) by comparing the trajectories of the QDs. The average dwell times were consistent with known kinetics of the actomyosin system, and the distribution of the waiting time between observed events was consistent with a Poisson process and the expected ATPase rate. Geometric constraints suggest a maximum of ∼26 nm of S2 can be unzipped from the filament, presumably involving disruption in the coiled-coil S2, a result consistent with observations by others of S2 protruding from the filament in muscle. We propose that sufficient force is available from the working heads in the filament to overcome the stiffness imposed by filament-S2 interactions.  相似文献   

19.
We report on the effects of self-assembled monolayer (SAM) dilution and thickness on the electron transfer (ET) event for cytochrome c (CytC) electrostatically immobilized on carboxyl terminated groups. We observed biphasic kinetic behavior for a logarithmic dependence of the rate constant on the SAM carbon number (ET distance) within the series of mixed SAMs of C(5)COOH/C(2)OH, C(10)COOH/C(6)OH, and C(15)COOH/C(11)OH that is in overall similar to that found earlier for the undiluted SAM assemblies. However, in the case of C(15)COOH/C(11)OH and C(10)COOH/C(6)OH mixed SAMs a notable increase of the ET standard rate constant was observed, in comparison with the corresponding unicomponent (omega-COOH) SAMs. In the case of the C(5)COOH/C(2)OH composite SAM a decrease of the rate constant versus the unicomponent analogue was observed. The value of the reorganization free energy deduced through the Marcus-like data analysis did not change throughout the series; this fact along with the other observations indicates uncomplicated rate-determining unimolecular ET in all cases. Our results are consistent with a model that considers a changeover between the alternate, tunneling and adiabatic intrinsic ET mechanisms. The physical mechanism behind the observed fine kinetic effects in terms of the protein-rigidifying omega-COOH/CytC interactions arising in the case of mixed SAMs are also discussed.  相似文献   

20.
RNA can play dual roles as a carrier of genetic information and as a catalyst of specific reactions, and it may have been the first biopolymer to have emerged on the early earth. The non-enzymatic replication of RNA was likely a key step in the evolution of simple cellular life from prebiotic chemistry. In the current model of template-directed polymerization of activated monomers, the chemical copying of RNA always generates a mixture of 3′–5′ and 2′–5′ backbone linkages due to the similar nucleophilicity and orientation of the 2′ and 3′ hydroxyl groups on the ribose. This lack of regiospecificity has been regarded as a central problem for the evolution of functional RNAs, since the resulting backbone heterogeneity was expected to disrupt their folding, molecular recognition and catalytic properties of functional RNAs such as ribozymes. However, a recent study from our lab has demonstrated that RNAs with a certain percentage of 2′–5′ linkages can still retain RNA functions, for example, in a FMN-binding aptamer and a hammerhead ribozyme system. More interestingly, it has been known for a long time that 2′–5′ linkages can reduce the melting temperature of RNA duplexes, making it easier to separate the strands. Although the detailed mechanism is still not clear, considering that strand separation is another unsolved big problem for non-enzymatic RNA replication, this feature may actually afford a selective advantage to duplexes exhibiting backbone heterogeneity. In addition, previous studies have revealed that 2′–5′ linkages in a RNA duplex are more easily hydrolyzed compared to normal 3′–5′ linkages. Thus, there is a selective advantage for the evolution of homogeneous RNA systems with more accurate replication. Altogether, the coexistence of 2′–5′ and 3′–5′ linkages may be a central feature that allowed RNA to play a central role in the original stage of life. In this work, we will present several X-ray crystal structures of RNA duplexes and an aptamer that contain 2′–5′ linkages. These structures help us to understand how RNA can adjust its structure to accommodate the backbone heterogeneity.  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