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 187 毫秒
1.
ATP-dependent DNA ligases catalyze the sealing of 5′-phosphate and 3′-hydroxyl termini at DNA nicks by means of a series of three nucleotidyl transfer steps. Here we have analyzed by site-directed mutagenesis the roles of conserved amino acids of Chlorella virus DNA ligase during the third step of the ligation pathway, which entails reaction of the 3′-OH of the nick with the DNA–adenylate intermediate to form a phosphodiester and release AMP. We found that Asp65 and Glu67 in nucleotidyltransferase motif III and Glu161 in motif IV enhance the rate of step 3 phosphodiester formation by factors of 20, 1000 and 60, respectively. Asp29 and Arg32 in nucleotidyltransferase motif I enhance the rate of step 3 by 60-fold. Gel shift analysis showed that mutations of Arg32 and Asp65 suppressed ligase binding to a pre-adenylated nick, whereas Asp29, Glu67 and Glu161 mutants bound stably to DNA–adenylate. We infer that Asp29, Glu67 and Glu161 are involved directly in the step 3 reaction. In several cases, the effects of alanine or conservative mutations on step 3 were modest compared to their effects on the composite ligation reaction and individual upstream steps. These results, in concert with available crystallographic data, suggest that the active site of DNA ligase is remodeled during the three steps of the pathway and that some of the catalytic side chains play distinct roles at different stages.  相似文献   

2.
Chlorella virus DNA ligase is the smallest eukaryotic ATP-dependent DNA ligase known; it suffices for yeast cell growth in lieu of the essential yeast DNA ligase Cdc9. The Chlorella virus ligase–adenylate intermediate has an intrinsic nick sensing function and its DNA footprint extends 8–9 nt on the 3′-hydroxyl (3′-OH) side of the nick and 11–12 nt on the 5′-phosphate (5′-PO4) side. Here we establish the minimal length requirements for ligatable 3′-OH and 5′-PO4 strands at the nick (6 nt) and describe a new crystal structure of the ligase–adenylate in a state construed to reflect the configuration of the active site prior to nick recognition. Comparison with a previous structure of the ligase–adenylate bound to sulfate (a mimetic of the nick 5′-PO4) suggests how the positions and contacts of the active site components and the bound adenylate are remodeled by DNA binding. We find that the minimal Chlorella virus ligase is capable of catalyzing non-homologous end-joining reactions in vivo in yeast, a process normally executed by the structurally more complex cellular Lig4 enzyme. Our results suggest a model of ligase evolution in which: (i) a small ‘pluripotent’ ligase is the progenitor of the much larger ligases found presently in eukaryotic cells and (ii) gene duplications, variations within the core ligase structure and the fusion of new domains to the core structure (affording new protein–protein interactions) led to the compartmentalization of eukaryotic ligase function, i.e. by enhancing some components of the functional repertoire of the ancestral ligase while disabling others.  相似文献   

3.
An NAD+-dependent DNA ligase from the hyperthermophilic bacterium Aquifex aeolicus was cloned, expressed in Escherichia coli and purified to homogeneity. The enzyme is most active in slightly alkaline pH conditions with either Mg2+ or Mn2+ as the metal cofactor. Ca2+ and Ni2+ mainly support formation of DNA–adenylate intermediates. The catalytic cycle is characterized by a low kcat value of 2 min–1 with concomitant accumulation of the DNAadenylate intermediate when Mg2+ is used as the metal cofactor. The ligation rates of matched substrates vary by up to 4-fold, but exhibit a general trend of T/A G/C < C/G < A/T on both the 3′- and 5′-side of the nick. Consistent with previous studies on Thermus ligases, this Aquifex ligase exhibits greater discrimination against a mismatched base pair on the 3′-side of the nick junction. The requirement of 3′ complementarity for a ligation reaction is reaffirmed by results from 1 nt insertions on either the 3′- or 5′-side of the nick. Furthermore, most of the unligatable 3′ mismatched base pairs prohibit formation of the DNAadenylate intermediate, indicating that the substrate adenylation step is also a control point for ligation fidelity. Unlike previously studied ATP ligases, gapped substrates cannot be ligated and intermediate accumulation is minimal, suggesting that complete elimination of base pair complementarity on one side of the nick affects substrate adenylation on the 5′-side of the nick junction. Relationships among metal cofactors, ligation products and intermediate, and ligation fidelity are discussed.  相似文献   

4.
The proteome of the amoebo-flagellate protozoan Naegleria gruberi is rich in candidate RNA repair enzymes, including 15 putative RNA ligases, one of which, NgrRnl, is a eukaryal homolog of Deinococcus radiodurans RNA ligase, DraRnl. Here we report that purified recombinant NgrRnl seals nicked 3′-OH/5′-PO4 duplexes in which the 3′-OH strand is RNA. It does so via the “classic” ligase pathway, entailing reaction with ATP to form a covalent NgrRnl–AMP intermediate, transfer of AMP to the nick 5′-PO4, and attack of the RNA 3′-OH on the adenylylated nick to form a 3′–5′ phosphodiester. Unlike members of the four known families of ATP-dependent RNA ligases, NgrRnl lacks a carboxy-terminal appendage to its nucleotidyltransferase domain. Instead, it contains a defining amino-terminal domain that we show is important for 3′-OH/5′-PO4 nick-sealing and ligase adenylylation, but dispensable for phosphodiester synthesis at a preadenylylated nick. We propose that NgrRnl, DraRnl, and their homologs from diverse bacteria, viruses, and unicellular eukarya comprise a new “Rnl5 family” of nick-sealing ligases with a signature domain organization.  相似文献   

5.
Escherichia coli DNA ligase (EcoLigA) repairs 3′-OH/5′-PO4 nicks in duplex DNA via reaction of LigA with NAD+ to form a covalent LigA-(lysyl-Nζ)–AMP intermediate (step 1); transfer of AMP to the nick 5′-PO4 to form an AppDNA intermediate (step 2); and attack of the nick 3′-OH on AppDNA to form a 3′-5′ phosphodiester (step 3). A distinctive feature of EcoLigA is its stimulation by ammonium ion. Here we used rapid mix-quench methods to analyze the kinetic mechanism of single-turnover nick sealing by EcoLigA–AMP. For substrates with correctly base-paired 3′-OH/5′-PO4 nicks, kstep2 was fast (6.8–27 s−1) and similar to kstep3 (8.3–42 s−1). Absent ammonium, kstep2 and kstep3 were 48-fold and 16-fold slower, respectively. EcoLigA was exquisitely sensitive to 3′-OH base mispairs and 3′ N:abasic lesions, which elicited 1000- to >20000-fold decrements in kstep2. The exception was the non-canonical 3′ A:oxoG configuration, which EcoLigA accepted as correctly paired for rapid sealing. These results underscore: (i) how EcoLigA requires proper positioning of the nick 3′ nucleoside for catalysis of 5′ adenylylation; and (ii) EcoLigA''s potential to embed mutations during the repair of oxidative damage. EcoLigA was relatively tolerant of 5′-phosphate base mispairs and 5′ N:abasic lesions.  相似文献   

6.
T4 RNA ligase 2 (Rnl2) repairs 3′-OH/5′-PO4 nicks in duplex nucleic acids in which the broken 3′-OH strand is RNA. Ligation entails three chemical steps: reaction of Rnl2 with ATP to form a covalent Rnl2–(lysyl-Nζ)–AMP intermediate (step 1); transfer of AMP to the 5′-PO4 of the nick to form an activated AppN– intermediate (step 2); and attack by the nick 3′-OH on the AppN– strand to form a 3′–5′ phosphodiester (step 3). Here we used rapid mix-quench methods to analyze the kinetic mechanism and fidelity of single-turnover nick sealing by Rnl2–AMP. For substrates with correctly base-paired 3′-OH nick termini, kstep2 was fast (9.5 to 17.9 sec−1) and similar in magnitude to kstep3 (7.9 to 32 sec−1). Rnl2 fidelity was enforced mainly at the level of step 2 catalysis, whereby 3′-OH base mispairs and oxoguanine, oxoadenine, or abasic lesions opposite the nick 3′-OH elicited severe decrements in the rate of 5′-adenylylation and relatively modest slowing of the rate of phosphodiester synthesis. The exception was the noncanonical A:oxoG base pair, which Rnl2 accepted as a correctly paired end for rapid sealing. These results underscore (1) how Rnl2 requires proper positioning of the 3′-terminal ribonucleoside at the nick for optimal 5′-adenylylation and (2) the potential for nick-sealing ligases to embed mutations during the repair of oxidative damage.  相似文献   

7.
Chlorella virus PBCV-1 DNA ligase seals nicked DNA substrates consisting of a 5'-phosphate-terminated strand and a 3'-hydroxyl-terminated strand annealed to a bridging DNA template strand. The enzyme discriminates at the DNA binding step between substrates containing a 5'-phosphate versus a 5'-hydroxyl at the nick. Mutational analysis of the active site motif KxDGxR (residues 27-32) illuminates essential roles for the conserved Lys, Asp and Arg moieties at different steps of the ligase reaction. Mutant K27A is unable to form the covalent ligase-(Lys-straightepsilonN-P)-adenylate intermediate and hence cannot activate a nicked DNA substrate via formation of the DNA-adenylate intermediate. Nonetheless, K27A catalyzes phosphodiester bond formation at a pre-adenylated nick. This shows that the active site lysine is not required for the strand closure reaction. K27A binds to nicked DNA-adenylate, but not to a standard DNA nick. This suggests that occupancy of the AMP binding pocket of DNA ligase is important for nick recognition. Mutant D29A is active in enzyme-adenylate formation and binds readily to nicked DNA, but is inert in DNA-adenylate formation. R32A is unable to catalyze any of the three reactions of the ligation pathway and does not bind to nicked DNA.  相似文献   

8.
9.

Background

RNA ligases are essential reagents for many methods in molecular biology including NextGen RNA sequencing. To prevent ligation of RNA to itself, ATP independent mutant ligases, defective in self-adenylation, are often used in combination with activated pre-adenylated linkers. It is important that these ligases not have de-adenylation activity, which can result in activation of RNA and formation of background ligation products. An additional useful feature is for the ligase to be active at elevated temperatures. This has the advantage or reducing preferences caused by structures of single-stranded substrates and linkers.

Results

To create an RNA ligase with these desirable properties we performed mutational analysis of the archaeal thermophilic RNA ligase from Methanobacterium thermoautotrophicum. We identified amino acids essential for ATP binding and reactivity but dispensable for phosphodiester bond formation with 5’ pre-adenylated donor substrate. The motif V lysine mutant (K246A) showed reduced activity in the first two steps of ligation reaction. The mutant has full ligation activity with pre-adenylated substrates but retained the undesirable activity of deadenylation, which is the reverse of step 2 adenylation. A second mutant, an alanine substitution for the catalytic lysine in motif I (K97A) abolished activity in the first two steps of the ligation reaction, but preserved wild type ligation activity in step 3. The activity of the K97A mutant is similar with either pre-adenylated RNA or single-stranded DNA (ssDNA) as donor substrates but we observed two-fold preference for RNA as an acceptor substrate compared to ssDNA with an identical sequence. In contrast, truncated T4 RNA ligase 2, the commercial enzyme used in these applications, is significantly more active using pre-adenylated RNA as a donor compared to pre-adenylated ssDNA. However, the T4 RNA ligases are ineffective in ligating ssDNA acceptors.

Conclusions

Mutational analysis of the heat stable RNA ligase from Methanobacterium thermoautotrophicum resulted in the creation of an ATP independent ligase. The K97A mutant is defective in the first two steps of ligation but retains full activity in ligation of either RNA or ssDNA to a pre-adenylated linker. The ability of the ligase to function at 65°C should reduce the constraints of RNA secondary structure in RNA ligation experiments.  相似文献   

10.
Deinococcus radiodurans RNA ligase (DraRnl) is a template-directed ligase that seals nicked duplexes in which the 3′-OH strand is RNA. DraRnl is a 342 amino acid polypeptide composed of a C-terminal adenylyltransferase domain fused to a distinctive 126 amino acid N-terminal module (a putative OB-fold). An alanine scan of the C domain identified 9 amino acids essential for nick ligation, which are located within nucleotidyltransferase motifs I, Ia, III, IIIa, IV and V. Seven mutants were dysfunctional by virtue of defects in ligase adenylylation: T163A, H167A, G168A, K186A, E230A, F281A and E305A. Four of these were also defective in phosphodiester formation at a preadenylylated nick: G168A, E230A, F281A and E305A. Two nick sealing-defective mutants were active in ligase adenylylation and sealing a preadenylylated nick, thereby implicating Ser185 and Lys326 in transfer of AMP from the enzyme to the nick 5′-PO4. Whereas deletion of the N-terminal domain suppressed overall nick ligation and ligase adenylylation, it did not compromise sealing at a preadenylylated nick. Mutational analysis of 15 residues of the N domain identified Lys26, Gln31 and Arg79 as key constituents. Structure–activity relationships at the essential residues were determined via conservative substitutions. We propose that DraRnl typifies a new clade of polynucleotide ligases. DraRnl homologs are detected in several eukaryal proteomes.  相似文献   

11.
A conserved catalytic core of the ATP-dependent DNA ligases is composed of an N-terminal domain (domain 1, containing nucleotidyl transferase motifs I, III, IIIa and IV) and a C-terminal domain (domain 2, containing motif VI) with an intervening cleft. Motif V links the two structural domains. Deletion analysis of the 298 amino acid Chlorella virus DNA ligase indicates that motif VI plays a critical role in the reaction of ligase with ATP to form ligase-adenylate, but is dispensable for the two subsequent steps in the ligation pathway; DNA-adenylate formation and strand closure. We find that formation of a phosphodiester at a pre-adenylated nick is subject to a rate limiting step that does not apply during the sealing of nicked DNA by ligase-adenylate. This step, presumably conformational, is accelerated or circumvented by deleting five amino acids of motif VI. The motif I lysine nucleophile (Lys27) is not required for strand closure by wild-type ligase, but this residue enhances the closure rate by a factor of 16 when motif VI is truncated. We find that a more extensively truncated ligase consisting of only N-terminal domain 1 and motif V is inert in ligase--adenylate formation, but competent to catalyze strand closure at a pre-adenylated nick. These results suggest that different enzymic catalysts facilitate the three steps of the DNA ligase reaction.  相似文献   

12.
RNA ligases participate in repair, splicing and editing pathways that either reseal broken RNAs or alter their primary structure. Here, we report the characterization of an RNA ligase from the thermophilic archaeon, Methanobacterium thermoautotrophicum. The 381-amino acid Methanobacterium RNA ligase (MthRnl) catalyzes intramolecular ligation of 5′-PO4 single-strand RNA to form a covalently closed circular RNA molecule through ligase-adenylylate and RNA-adenylylate (AppRNA) intermediates. At the optimal temperature of 65°C, AppRNA was predominantly ligated to a circular product. In contrast, at 35°C, phosphodiester bond formation was suppressed and the majority of the AppRNA was deadenylylated. Sedimentation analysis indicates that MthRnl is a homodimer in solution. The C-terminal 127-amino acid segment is required for dimerization, is itself capable of oligomeization and acts in trans to inhibit the ligation activity of native MthRnl. MthRnl can also join single-stranded DNA to form a circular molecule. The lack of specificity for RNA and DNA by MthRnl may exemplify an undifferentiated ancestral stage in the evolution of ATP-dependent ligases.  相似文献   

13.
Deinococcus radiodurans RNA ligase (DraRnl) is the founding member of a family of end-joining enzymes encoded by diverse microbes and viruses. DraRnl ligates 3′-OH, 5′-PO4 nicks in double-stranded nucleic acids in which the nick 3′-OH end is RNA. Here we gauge the effects of 3′-OH and 5′-PO4 base mispairs and damaged base lesions on the rate of nick sealing. DraRnl is indifferent to the identity of the 3′-OH nucleobase, provided that it is correctly paired. With 3′-OH mispairs, the DraRnl sealing rate varies widely, with G-T and A-C mispairs being the best substrates and G-G, G-A, and A-A mispairs being the worst. DraRnl accepts 3′ A–8-oxoguanine (oxoG) to be correctly paired, while it discriminates against U-oxoG and G-oxoG mispairs. DraRnl displays high activity and low fidelity in sealing 3′-OH ends opposite an 8-oxoadenine lesion. It prefers 3′-OH adenosine when sealing opposite an abasic template site. With 5′-PO4 mispairs, DraRnl seals a 5′ T-G mispair as well as it does a 5′ C-G pair; in most other respects, the ligation fidelity at 5′ mispairs is similar to that at 3′ mispairs. DraRnl accepts a 5′ A-oxoG end to be correctly paired, yet it is more tolerant of 5′ T-oxoG and 5′ G-oxoG mispairs than the equivalent configurations on the 3′ side of the nick. At 5′ nucleobase-abasic site nicks, DraRnl prefers to ligate when the nucleobase is a purine. The biochemical properties of DraRnl are compatible with its participation in the templated repair of RNA damage or in the sealing of filled DNA gaps that have a 3′ ribopatch.  相似文献   

14.
Attachment of proteins to the 3′ end of DNA increases stability of the DNA in serum and retards clearance of DNA by major organs, thereby enhancing in vivo half-life and therapeutic potential of DNA. Unfortunately, the length of DNA molecules that can be produced with 3 ′ modifications by solid-phase synthesis for protein attachment is limited to 45–60 nucleotides due to uncertainties about sequence fidelity for longer oligonucleotides. Here we describe selective covalent coupling of proteins or other molecules to the 3′-adenine overhang of unlabeled and fluorophore-labeled double-stranded polymerase chain reaction products putatively at the N6 position of adenine using 2.5% glutaraldehyde at pH 6.0 and 4°C for at least 16 h. Gel mobility shift analyses and fluorescence analyses of the shifted bands supported conjugate formation between double-stranded polymerase chain reaction products and β2-microglobulin. In addition, blunt-ended DNA ladder fragments treated with glutaraldehyde at 4°C showed no evidence of DNA–DNA or DNA–protein conjugate formation. With the present cold glutaraldehyde technique, longer DNA–3′-protein conjugates might be easily mass-produced. The protein portion of a DNA–3′-protein conjugate could possess functionality as well, such as receptor binding for cell entry, cytotoxicity, or opsonization.  相似文献   

15.
2H (two-histidine) phosphoesterase enzymes are distributed widely in all domains of life and are implicated in diverse RNA and nucleotide transactions, including the transesterification and hydrolysis of cyclic phosphates. Here we report a biochemical and structural characterization of the Escherichia coli 2H protein YapD, which was identified originally as a reversible transesterifying “nuclease/ligase” at RNA 2′,5′-phosphodiesters. We find that YapD is an “end healing” cyclic phosphodiesterase (CPDase) enzyme that hydrolyzes an HORNA>p substrate with a 2′,3′-cyclic phosphodiester to a HORNAp product with a 2′-phosphomonoester terminus, without concomitant end joining. Thus we rename this enzyme ThpR (two-histidine 2′,3′-cyclic phosphodiesterase acting on RNA). The 2.0 Å crystal structure of ThpR in a product complex with 2′-AMP highlights the roles of extended histidine-containing motifs 43HxTxxF48 and 125HxTxxR130 in the CPDase reaction. His43-Nε makes a hydrogen bond with the ribose O3′ leaving group, thereby implicating His43 as a general acid catalyst. His125-Nε coordinates the O1P oxygen of the AMP 2′-phosphate (inferred from geometry to derive from the attacking water nucleophile), pointing to His125 as a general base catalyst. Arg130 makes bidentate contact with the AMP 2′-phosphate, suggesting a role in transition-state stabilization. Consistent with these inferences, changing His43, His125, or Arg130 to alanine effaced the CPDase activity of ThpR. Phe48 makes a π–π stack on the adenine nucleobase. Mutating Phe28 to alanine slowed the CPDase by an order of magnitude. The tertiary structure and extended active site motifs of ThpR are conserved in a subfamily of bacterial and archaeal 2H enzymes.  相似文献   

16.
Pseudomonas putida MPE exemplifies a novel clade of manganese-dependent single-strand DNA endonuclease within the binuclear metallophosphoesterase superfamily. MPE is encoded within a widely conserved DNA repair operon. Via structure-guided mutagenesis, we identify His113 and His81 as essential for DNA nuclease activity, albeit inessential for hydrolysis of bis-p-nitrophenylphosphate. We propose that His113 contacts the scissile phosphodiester and serves as a general acid catalyst to expel the OH leaving group of the product strand. We find that MPE cleaves the 3′ and 5′ single-strands of tailed duplex DNAs and that MPE can sense and incise duplexes at sites of short mismatch bulges and opposite a nick. We show that MPE is an ambidextrous phosphodiesterase capable of hydrolyzing the ssDNA backbone in either orientation to generate a mixture of 3′-OH and 3′-PO4 cleavage products. The directionality of phosphodiester hydrolysis is dictated by the orientation of the water nucleophile vis-à-vis the OH leaving group, which must be near apical for the reaction to proceed. We propose that the MPE active site and metal-bound water nucleophile are invariant and the enzyme can bind the ssDNA productively in opposite orientations.  相似文献   

17.
ATP-dependent DNA ligases, NAD(+)-dependent DNA ligases, and GTP-dependent RNA capping enzymes are members of a covalent nucleotidyl transferase superfamily defined by a common fold and a set of conserved peptide motifs. Here we examined the role of nucleotidyl transferase motif V ((184)LLKMKQFKDAEAT(196)) in the nick joining reaction of Chlorella virus DNA ligase, an exemplary ATP-dependent enzyme. We found that alanine substitutions at Lys(186), Lys(188), Asp(192), and Glu(194) reduced ligase specific activity by at least an order of magnitude, whereas substitutions at Lys(191) and Thr(196) were benign. The K186A, D192A, and E194A changes had no effect on the rate of single-turnover nick joining by preformed ligase-adenylate but affected subsequent rounds of nick joining at the ligase adenylation step. Conservative substitutions K186R, D192E, and E194D partially restored activity, whereas K186Q, D192N, and E194Q substitutions did not. Alanine mutation of Lys(188) elicited distinctive catalytic defects, whereby single-turnover nick joining by K188A-adenylate was slowed by an order of magnitude, and high levels of the DNA-adenylate intermediate accumulated. The rate of phosphodiester bond formation at a pre-adenylated nick (step 3 of the ligation pathway) was slowed by the K188A change. Replacement of Lys(188) by arginine reversed the step 3 arrest, whereas glutamine substitution was ineffective. Gel-shift analysis showed that the Lys(188) mutants bound stably to DNA-adenylate. We infer that Lys(188) is involved in the chemical step of phosphodiester bond formation.  相似文献   

18.
ABSTRACT: BACKGROUND: RNA ligases are essential reagents for many methods in molecular biology including NextGen RNA sequencing. To prevent ligation of RNA to itself, ATP independent mutant ligases, defective in self-adenylation, are often used in combination with activated pre-adenylated linkers. It is important that these ligases not have de-adenylation activity, which can result in activation of RNA and formation of background ligation products. An additional useful feature is for the ligase to be active at elevated temperatures. This has the advantage or reducing preferences caused by structures of single-stranded substrates and linkers. RESULTS: To create an RNA ligase with these desirable properties we performed mutational analysis of the archaeal thermophilic RNA ligase from Methanobacterium thermoautotrophicum. We identified amino acids essential for ATP binding and reactivity but dispensable for phosphodiester bond formation with 5' pre-adenylated donor substrate. The motif V lysine mutant (K246A) showed reduced activity in the first two steps of ligation reaction. The mutant has full ligation activity with pre-adenylated substrates but retained the undesirable activity of deadenylation, which is the reverse of step 2 adenylation. A second mutant, an alanine substitution for the catalytic lysine in motif I (K97A) abolished activity in the first two steps of the ligation reaction, but preserved wild type ligation activity in step 3. The activity of the K97A mutant is similar with either pre-adenylated RNA or single-stranded DNA (ssDNA) as donor substrates but we observed two-fold preference for RNA as an acceptor substrate compared to ssDNA with an identical sequence. In contrast, truncated T4 RNA ligase 2, the commercial enzyme used in these applications, is significantly more active using pre-adenylated RNA as a donor compared to pre-adenylated ssDNA. However, the T4 RNA ligases are ineffective in ligating ssDNA acceptors. CONCLUSIONS: Mutational analysis of the heat stable RNA ligase from Methanobacterium thermoautotrophicum resulted in the creation of an ATP independent ligase. The K97A mutant is defective in the first two steps of ligation but retains full activity in ligation of either RNA or ssDNA to a pre-adenylated linker. The ability of the ligase to function at 65 deg C should reduce the constraints of RNA secondary structure in RNA ligation experiments.  相似文献   

19.
RNA and DNA ligases catalyze the formation of a phosphodiester bond between the 5′-phosphate and 3′-hydroxyl ends of nucleic acids. In this work, we describe the ability of the thermophilic RNA ligase MthRnl from Methanobacterium thermoautotrophicum to recognize and modify the 3′-terminal phosphate of RNA and single-stranded DNA (ssDNA). This ligase can use an RNA 3′p substrate to generate an RNA 2′,3′-cyclic phosphate or convert DNA3′p to ssDNA3′pp5′A. An RNA ligase from the Thermus scotoductus bacteriophage TS2126 and a predicted T4 Rnl1-like protein from Thermovibrio ammonificans, TVa, were also able to adenylate ssDNA 3′p. These modifications of RNA and DNA 3′-phosphates are similar to the activities of RtcA, an RNA 3′-phosphate cyclase. The initial step involves adenylation of the enzyme by ATP, which is then transferred to either RNA 3′p or DNA 3′p to generate the adenylated intermediate. For RNA 3′pp5′A, the third step involves attack of the adjacent 2′ hydroxyl to generate the RNA 2′,3′-cyclic phosphate. These steps are analogous to those in classical 5′ phosphate ligation. MthRnl and TS2126 RNA ligases were not able to modify a 3′p in nicked double-stranded DNA. However, T4 DNA ligase and RtcA can use 3′-phosphorylated nicks in double-stranded DNA to produce a 3′-adenylated product. These 3′-terminal phosphate-adenylated intermediates are substrates for deadenylation by yeast 5′Deadenylase. Our findings that classic ligases can duplicate the adenylation and phosphate cyclization activity of RtcA suggests that they have an essential role in metabolism of nucleic acids with 3′-terminal phosphates.  相似文献   

20.
Delta helicase is a 5′ to 3′ DNA helicase that partially co-purifies with DNA polymerase delta (pol delta) from fetal bovine thymus tissue. We describe the resolution of delta helicase from pol delta on heparin–agarose chromatography and its purification to apparent homogeneity by affinity purification on single-stranded DNA–cellulose chromatography, unique-sequence RNA–agarose chromatography, and ceramic hydroxyapatite chromatography. Delta helicase isolated from fetal bovine thymus had an apparent Mr of 115 kDa in SDS–PAGE, and photo-crosslinked to [α-32P]ATP. Tandem mass spectrometry peptide mass data derived from the bovine polypeptide matched to human UPF1 (HUPF1), a 5′ to 3′ RNA and DNA helicase, and a requisite component of the mRNA surveillance complex. Antisera against HUPF1 cross-reacted with delta helicase on western analysis, and delta helicase activity was immunoinactivated by pre-incubation with antibodies to HUPF1, suggesting that delta helicase is the bovine homolog of HUPF1. Immunoprecipitation experiments demonstrated that HUPF1 interacts with the 66-kDa third subunit of pol delta in vivo.  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