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.
A Tetrahymena intron nucleotide connected to the GTP/arginine site.   总被引:2,自引:1,他引:1       下载免费PDF全文
M Yarus  J Levine  G B Morin    T R Cech 《Nucleic acids research》1989,17(17):6969-6981
  相似文献   

2.
Self-splicing group I introns use guanosine as a nucleophile to cleave the 5' splice site. The guanosine-binding site has been localized to the G264-C311 base pair of the Tetrahymena intron on the basis of analysis of mutations that change the specificity of the nucleophile from G (guanosine) to 2AP (2-aminopurine ribonucleoside) (F. Michel et al. (1989) Nature 342, 391-395). We studied the effect of these mutations (G-U, A-C and A-U replacing G264-C311) in the L-21 ScaI version of the Tetrahymena ribozyme. In this enzymatic system (kcat/Km)G monitors the cleavage step. This kinetic parameter decreased by at least 5 x 10(3) when the G264-C311 base pair was mutated to an A-U pair, while (kcat/Km)2AP increased at least 40-fold. This amounted to an overall switch in specificity of at least 2 x 10(5). The nucleophile specificity (G > 2AP for the G-C and G-U pairs, 2AP > G for the A-U and A-C pairs) was consistent with the proposed hydrogen bond between the nucleotide at position 264 and N1 of the nucleophile. Unexpectedly, the A-U and A-C mutants showed a decrease of an order of magnitude in the rate of ribozyme-catalyzed hydrolysis of RNA, in which H2O or OH- replaces G as the nucleophile, whereas the G-U mutant showed a decrease of only 2-fold. The low hydrolysis rates were not restored by raising the Mg2+ concentration or lowering the temperature. In addition, the mutant ribozymes exhibited a pattern of cleavage by Fe(II)-EDTA indistinguishable from that of the wild type, and the [Mg2+]1/2 for folding of the A-U mutant ribozyme was the same as that of the wild type. Therefore the guanosine-binding site mutations do not appear to have a major effect on RNA folding or stability. Because changing G264 affects the hydrolysis reaction without perturbing the global folding of the RNA, we conclude that the catalytic role of this conserved nucleotide is not limited to guanosine binding.  相似文献   

3.
Identification of two novel arginine binding DNAs.   总被引:5,自引:0,他引:5       下载免费PDF全文
K Harada  A D Frankel 《The EMBO journal》1995,14(23):5798-5811
RNA tertiary structure is known to play critical roles in RNA-protein recognition and RNA function. To examine how DNA tertiary structure might relate to RNA structure, we performed in vitro selection experiments to identify single-stranded DNAs that specifically bind arginine, and compared the results with analogous experiments performed with RNA. In the case of RNA, a motif related to the arginine binding site in human immunodeficiency virus TAR RNA was commonly found, whereas in the case of DNA, two novel motifs and no TAR-like structures were found. One DNA motif, found in approximately 40% of the cloned sequences, forms of hairpin structure with a highly conserved 10 nucleotide loop, whereas the second motif is especially rich in G residues. Chemical interference and mutagenesis experiments identified nucleotides in both motifs that form specific arginine binding sites, and dimethylsulfate footprinting experiments identified single guanine residues in both that are protected from methylation in the presence of arginine, suggesting possible sites of arginine contact or conformational changes in the DNAs. Circular dichroism experiments indicated that both DNAs undergo conformational changes upon arginine binding and that the arginine guanidinium group alone is responsible for binding. A model for the G-rich motif is proposed in which mixed guanine and adenine quartets may form a novel DNA structure. Arginine binding DNAs and RNAs should provide useful model systems for studying nucleic acid tertiary structure.  相似文献   

4.
An RNA-amino acid complex and the origin of the genetic code   总被引:4,自引:0,他引:4  
M Yarus 《The New biologist》1991,3(2):183-189
The group I RNAs, of which the Tetrahymena ribosomal RNA intron is the most investigated example, catalyze their own splicing reactions. Splicing is initiated at a conserved site on the RNA that facilitates attack by exogenous guanosine (or its nucleotides) on the exon-intron junction. The guanosine site in the RNA's catalytic center also binds arginine, and is quite selective for the arginine side chain. This amino acid-RNA interaction is stereoselective, and L-arginine is preferred. Immediately at the site at which arginine binds there is one of only four RNA triplets in 92 group I RNA sequences: AGA/G and CGA/G. Thus the arginine contact site is within any of four different codons for arginine. Mutation of the conserved G in the middle of the triplet decreases affinity for the amino acid, showing that binding is sequence-specific. A pathway for the origin of the genetic code for arginine is suggested, based on the existence and properties of this sequence-specific, amino acid-specific RNA complex. The existence of a proto-ribosome related to the group I RNAs seems the most likely hypothesis. This notion is used to distinguish three periods in the development of the code. Restrained and exuberant hypotheses about the origin of the genetic code are distinguished, and some objections to these hypotheses are considered.  相似文献   

5.
6.
The beta subunit isolated from the chloroplast ATP synthase F1 (CF1) has a single dissociable nucleotide binding site, consistent with the proposed function of this subunit in nucleotide binding and catalysis. The beta subunit bound the nucleotide analogs trinitrophenyl-ATP (TNP-ATP) or trinitrophenyl-ADP (TNP-ADP) with nearly equal affinities (Kd = 1-2 microM) but did not bind trinitrophenyl-AMP. Both ATP and ADP effectively competed with TNP-ATP for binding. Other nucleoside triphosphates were also able to compete with TNP-ATP for binding to beta; their order of effectiveness (ATP greater than GTP, ITP greater than CTP) mimicked the normal substrate specificity of CF1. The single nucleotide binding site on the isolated beta subunit very closely resembles the low affinity catalytic site (site 3) of CF1 (Bruist, M.F., and Hammes, G. G. (1981) Biochemistry 20, 6298-6305), suggesting that tight nucleotide binding by other sites on the enzyme involves other CF1 subunits in addition to the beta subunit. The results are inconsistent with an earlier report (Frasch, W.D., Green, J., Caguial, J., and Mejia, A. (1989) J. Biol. Chem. 264, 5064-5069), which suggested more than one nucleotide binding site per beta subunit. Binding of nucleotides to the isolated beta subunit was eliminated by a brief heat treatment (40 degrees C for 10 min) of the protein. A small change in the circular dichroism spectrum of beta accompanied the heat treatment indicating that a localized (rather than global) change in the folding of beta, involving at least part of the nucleotide binding domain, had occurred. Also accompanying the loss of nucleotide binding was a loss of the reconstitutive capacity of the beta subunit. ATP protected against the effects of the heat treatment.  相似文献   

7.
An axial binding site in the Tetrahymena precursor RNA.   总被引:7,自引:0,他引:7  
Previous studies allow the construction of three distinct models of the binding of G and arginine within the active site of the Tetrahymena self-splicing preribosomal precursor RNA. These models (base triple, axial I and axial II) are now distinguished by measurements on the specificity of RNAs with nucleotide substitutions at positions spanning the site. Because the semi-conserved unpaired nucleotide 263 has no effect on substrate or inhibitor selection by the Tetrahymena RNA we conclude that the axial I model is improbable. In contrast, data with substituted RNAs and nucleoside analogs suggest that nucleotide 265 makes a hydrogen bond with the substrate. Accordingly the active site appears axial because substrate contacts exist at more than one nucleotide on the 5' side of the P7 helix. The effects of this hydrogen bond are observable in cases where the donor or acceptor is on the RNA, whether nucleotide 265 is a purine or pyrimidine, or whether nucleotide 265 is mispaired, wobble paired or normally paired. This pattern is consistent with the axial II model. Molecular dynamics and energy minimization calculations lead to the same conclusions as these site-directed substitutions; the base triple and axial I models are unstable dynamically. Under thermal agitation, the third model site (axial II) is transformed to a related, but more stable structure, axial III. The axial III active site is characterized by the extrusion of the conserved bulged base 263 from the P7 helix, a semi-pocket for G base formed by stacking of nucleotide 262, and formation of all bonds to the G base originally proposed for both the base triple and axial II sites. Because of these hydrogen bonds the axial III site is also consistent with data on enzymatic specificity. The axial III model indicates an unforeseen capacity for pocket formation within the groove of an RNA helix, suggests that the site may be unusually flexible, and bears on a hypothesis concerning the origin of the genetic code.  相似文献   

8.
Cai Z  Liang TJ  Luo G 《Journal of virology》2004,78(7):3633-3643
Replication of nearly all RNA viruses depends on a virus-encoded RNA-dependent RNA polymerase (RdRp). Our earlier work found that purified recombinant hepatitis C virus (HCV) RdRp (NS5B) was able to initiate RNA synthesis de novo by using purine (A and G) but not pyrimidine (C and U) nucleotides (G. Luo et al., J. Virol. 74:851-863, 2000). For most human RNA viruses, the initiation nucleotides of both positive- and negative-strand RNAs were found to be either an adenylate (A) or guanylate (G). To determine the nucleotide used for initiation and control of HCV RNA replication, a genetic mutagenesis analysis of the nucleotides at the very 5' and 3' ends of HCV RNAs was performed by using a cell-based HCV replicon replication system. Either a G or an A at the 5' end of HCV genomic RNA was able to efficiently induce cell colony formation, whereas a nucleotide C at the 5' end dramatically reduced the efficiency of cell colony formation. Likewise, the 3'-end nucleotide U-to-C mutation did not significantly affect the efficiency of cell colony formation. In contrast, a U-to-G mutation at the 3' end caused a remarkable decrease in cell colony formation, and a U-to-A mutation resulted in a complete abolition of cell colony formation. Sequence analysis of the HCV replicon RNAs recovered from G418-resistant Huh7 cells revealed several interesting findings. First, the 5'-end nucleotide G of the replicon RNA was changed to an A upon multiple rounds of replication. Second, the nucleotide A at the 5' end was stably maintained among all replicon RNAs isolated from Huh7 cells transfected with an RNA with a 5'-end A. Third, initiation of HCV RNA replication with a CTP resulted in a >10-fold reduction in the levels of HCV RNAs, suggesting that initiation of RNA replication with CTP was very inefficient. Fourth, the 3'-end nucleotide U-to-C and -G mutations were all reverted back to a wild-type nucleotide U. In addition, extra U and UU residues were identified at the 3' ends of revertants recovered from Huh7 cells transfected with an RNA with a nucleotide G at the 3' end. We also determined the 5'-end nucleotide of positive-strand RNA of some clinical HCV isolates. Either G or A was identified at the 5' end of HCV RNA genome depending on the specific HCV isolate. Collectively, these findings demonstrate that replication of positive-strand HCV RNA was preferentially initiated with purine nucleotides (ATP and GTP), whereas the negative-strand HCV RNA replication is invariably initiated with an ATP.  相似文献   

9.
10.
mRNA-specific polyadenylation can be assayed in vitro by using synthetic RNAs that end at or near the natural cleavage site. This reaction requires the highly conserved sequence AAUAAA. At least two distinct nuclear components, an AAUAAA specificity factor and poly(A) polymerase, are required to catalyze the reaction. In this study, we identified structural features of the RNA substrate that are critical for mRNA-specific polyadenylation. We found that a substrate that contained only 11 nucleotides, of which the first six were AAUAAA, underwent AAUAAA-specific polyadenylation. This is the shortest substrate we have used that supports polyadenylation: removal of a single nucleotide from either end of this RNA abolished the reaction. Although AAUAAA appeared to be the only strict sequence requirement for polyadenylation, the number of nucleotides between AAUAAA and the 3' end was critical. Substrates with seven or fewer nucleotides beyond AAUAAA received poly(A) with decreased efficiency yet still bound efficiently to specificity factor. We infer that on these shortened substrates, poly(A) polymerase cannot simultaneously contact the specificity factor bound to AAUAAA and the 3' end of the RNA. By incorporating 2'-deoxyuridine into the U of AAUAAA, we demonstrated that the 2' hydroxyl of the U in AAUAAA was required for the binding of specificity factor to the substrate and hence for poly(A) addition. This finding may indicate that at least one of the factors involved in the interaction with AAUAAA is a protein.  相似文献   

11.
12.
M S Rohrbach  J W Bodley 《Biochemistry》1977,16(7):1360-1363
Treatment of Escherichia coli elongation factor G with the arginine reagent, 2,3-butanedione, leads to the inactivation of the enzyme when performed in sodium borate buffers. The inhibition follows pseudo-first-order kinetics until 95% of the activity has been lost and further incubation results in complete inhibiton. Removal of the borate by exhaustive dialysis results in the restoration of approximately 85% of the original activity. The pH dependence of the reaction suggests that the ionization of a group in the protein with a pKa of approximately 8.8 facilitates the reaction with butanedione. A reaction order of 1.01 +/- 0.13 was calculated for the inhibition reaction, indicating that the incorporation of one butanedione per elongation factor G results in the inactivation of the enzyme. The kinetics of inhibition in the presence of GTP indicate that the elongation factor G-GTP complex is refractory to butanedione inhibiton. Elongation factor G which has been partially inactivated by butanedione has the same apparent Km for GTP as does the native enzyme. These results indicate that elongation factor G contains only one essential arginine residue which is reactive with butanedione and that this residue is located at its nucleotide binding site.  相似文献   

13.
CTP synthase catalyzes the reaction glutamine + UTP + ATP --> glutamate + CTP + ADP + Pi. The rate of the reaction is greatly enhanced by the allosteric activator GTP. We have studied the glutaminase half-reaction of CTP synthase from Lactococcus lactis and its response to the allosteric activator GTP and nucleotides that bind to the active site. In contrast to what has been found for the Escherichia coli enzyme, GTP activation of the L. lactis enzyme did not result in similar kcat values for the glutaminase activity and glutamine hydrolysis coupled to CTP synthesis. GTP activation of the glutaminase reaction never reached the levels of GTP-activated CTP synthesis, not even when the active site was saturated with UTP and the nonhydrolyzeable ATP-binding analog adenosine 5'-[gamma-thio]triphosphate. Furthermore, under conditions where the rate of glutamine hydrolysis exceeded that of CTP synthesis, GTP would stimulate CTP synthesis. These results indicate that the L. lactis enzyme differs significantly from the E. coli enzyme. For the E. coli enzyme, activation by GTP was found to stimulate glutamine hydrolysis and CTP synthesis to the same extent, suggesting that the major function of GTP binding is to activate the chemical steps of glutamine hydrolysis. An alternative mechanism for the action of GTP on L. lactis CTP synthase is suggested. Here the binding of GTP to the allosteric site promotes coordination of the phosphorylation of UTP and hydrolysis of glutamine for optimal efficiency in CTP synthesis rather than just acting to increase the rate of glutamine hydrolysis itself.  相似文献   

14.
There are two functionally and physically distinct types of guanyl nucleotide site associated with the adenylate cyclase system of pigeon erythrocytes. One is on the well known regulatory protein, N, that mediates the adenylate cyclase response to hormones, guanyl nucleotides and fluoride, and is the substrate for ADP-ribosylation by cholera toxin. We now describe a second site that must be occupied by GTP or an analog of GTP before N can be ADP-ribosylated. We call this second site S. It differs from the site on N in many respects. GTP appears to be rapidly hydrolyzed when it is bound to N but not when bound at S. GTP analogs such as guanosine 5'-O-(3-thiotriphosphate) (GTP gamma S) bind stably to both sites but the binding of GTP gamma S to N is more sensitive to EDTA and is more easily prevented by guanosine 5'-O-(2-thiodiphosphate). The nucleotide binding only to S is promoted by the cytosolic protein required by cholera toxin. Isoproterenol decreases GTP gamma S binding to S while indirectly increasing GTP gamma S binding to N. By adjusting the binding conditions, the nucleotides bound functionally to N and S can be varied independently and then the effect of ADP-ribosylation upon the adenylate cyclase activity can be seen to depend on the type of nucleotide bound to N. This activity rises, falls slightly, or remains at zero, if N is occupied by GTP, GTP gamma S, or guanosine 5'-O-(2-thiodiphosphate, respectively.  相似文献   

15.
RNA molecules with internal 2′,5′-branches are intermediates in RNA splicing, and branched RNAs have recently been proposed as retrotransposition intermediates. A broadly applicable in vitro synthetic route to branched RNA that does not require self-splicing introns or spliceosomes would substantially improve our ability to study biochemical processes that involve branched RNA. We recently described 7S11, a deoxyribozyme that was identified by in vitro selection and has general RNA branch-forming ability. However, an important restriction for 7S11 is that the branch-site RNA nucleotide must be a purine (A or G), because a pyrimidine (U or C) is not tolerated. Here, we describe the compact 6CE8 deoxyribozyme (selected using a 20 nt random region) that synthesizes 2′,5′-branched RNA with any nucleotide at the branch site. The Mn2+-dependent branch-forming ligation reaction is between an internal branch-site 2′-hydroxyl nucleophile on one RNA substrate with a 5′-triphosphate on another RNA substrate. The preference for the branch-site nucleotide is U > C A > G, although all four nucleotides are tolerated with useful ligation rates. Nearly all other nucleotides elsewhere in both RNA substrates allow ligation activity, except that the sequence requirement for the RNA strand with the 5′-triphosphate is 5′-pppGA, with 5′-pppGAR (R = purine) preferred. These characteristics permit 6CE8 to prepare branched RNAs of immediate practical interest, such as the proposed branched intermediate of Ty1 retrotransposition. Because this branched RNA has two strands with identical sequence that emerge from the branch site, we developed strategies to control which of the two strands bind with the deoxyribozyme during the branch-forming reaction. The ability to synthesize the proposed branched RNA of Ty1 retrotransposition will allow us to explore this important biochemical pathway in greater detail.  相似文献   

16.
Influenza virus uses a unique cap-snatching mechanism characterized by hijacking and cleavage of host capped pre-mRNAs, resulting in short capped RNAs, which are used as primers for viral mRNA synthesis. The PA subunit of influenza polymerase carries the endonuclease activity that catalyzes the host mRNA cleavage reaction. Here, we show that PA is a sequence selective endonuclease with distinct preference to cleave at the 3′ end of a guanine (G) base in RNA. The G specificity is exhibited by the native influenza polymerase complex associated with viral ribonucleoprotein particles and is conferred by an intrinsic G specificity of the isolated PA endonuclease domain PA-Nter. In addition, RNA cleavage site choice by the full polymerase is also guided by cap binding to the PB2 subunit, from which RNA cleavage preferentially occurs at the 12th nt downstream of the cap. However, if a G residue is present in the region of 10–13 nucleotides from the cap, cleavage preferentially occurs at G. This is the first biochemical evidence of influenza polymerase PA showing intrinsic sequence selective endonuclease activity.  相似文献   

17.
RNA editing in flowering plant mitochondria addresses several hundred specific C nucleotides in individual sequence contexts in mRNAs and tRNAs. Many of the in vivo steady state RNAs are edited at some sites but not at others. It is still unclear whether such incompletely edited RNAs can either be completed or are aborted. To learn more about the dynamics of the substrate recognition process, we investigated in vitro RNA editing at a locus in the atp4 mRNA where three editing sites are clustered within four nucleotides. A single cis-element of about 20 nucleotides serves in the recognition of at least two sites. Competition with this sequence element suppresses in vitro editing. Surprisingly, unedited and edited competitors are equally effective. Experiments with partially pre-edited substrates indicate that indeed the editing status of a substrate RNA does not affect the binding affinity of the specificity factor(s). RNA molecules in which all editing sites are substituted by either A or G still compete, confirming that editing site recognition can occur independently of the actual editing site. These results show that incompletely edited mRNAs can be substrates for further rounds of RNA editing, resolving a long debated question.  相似文献   

18.
Stoichiometric exchange of GTP for GDP on heterotrimeric G protein alpha (Galpha) subunits is essential to most hormone and neurotransmitter initiated signal transduction. Galphas are stably activated in a Mg2+ complex with GTPgammaS, a nonhydrolyzable GTP analogue that is reported to bind Galpha, with very high affinity. Yet, it is common to find that substantial amounts (30-90%) of purified G proteins cannot be activated. Inactivatable G protein has heretofore been thought to have become "denatured" during formation of the obligatory nucleotide-free or empty (MT) Galpha-state that is intermediary to GDP/GTP exchange at a single binding site. We find Galpha native secondary and tertiary structure to persist during formation of the irreversibly inactivatable state of transducin. MT Galpha is therefore irreversibly misfolded rather than denatured. Inactivation by misfolding is found to compete kinetically with protective but weak preequilibrium nucleotide binding at micromolar ambient GTPgammaS concentrations. Because of the weak preequilibrium, quantitative protection against Galpha aggregation is only achieved at free nucleotide concentrations 10-100 times higher than those commonly employed in G protein radio-nucleotide binding studies. Initial GTP protection is also poor because of the extreme slowness of an intramolecular Galpha refolding step (isomerization) necessary for GTP sequestration after its weak preequilibrium binding. Of the two slowly interconverting Galpha x GTP isomers described here, only the second can bind Mg2+, "locking" GTP in place with a large net rise in GTP binding affinity. A companion Galpha x GDP isomerization reaction is identified as the cause of the very slow spontaneous GDP dissociation that characterizes G protein nucleotide exchange and low spontaneous background activity in the absence of GPCR activation. Galpha x GDP and Galpha x GTP isomerization reactions are proposed as the dual target for GPCR catalysis of nucleotide exchange.  相似文献   

19.
The chick cerebellar kainate (KA) binding protein (KBP), a member of the family of ionotropic glutamate receptors, harbours a glycine-rich (GxGxxG) motif known to be involved in the binding of ATP and GTP to kinases and G proteins respectively. Here, we report that guanine, but not adenine, nucleotides interact with KBP by inhibiting [3H]KA binding in a competitive-like manner, displaying IC50 values in the micromolar range. To locate the GTP binding site, KBP was photoaffinity labelled with [alpha-32P]GTP. The reaction was blocked by KA, glutamate, 6-cyano-7-nitroquinoxaline-2,3-dione and antibodies raised against a peptide containing the glycine-rich motif. Site-directed mutagenesis of residues K72 and Y73 within the glycine-rich motif followed by the expression of the KBP mutants at the surface of HEK 293 cells showed a decrease in GTP binding affinity by factors of 10 and 100 respectively. The binding of [3H]KA to the K72A/T KBP mutants was not affected but binding to the Y73I KBP mutant was decreased by a factor of 10. Accordingly, we propose that the glycine-rich motif of KBP forms part of a guanine nucleotide binding site. We further suggest that the glycine-rich motif is the binding site at which guanine nucleotides inhibit the glutamate-mediated responses of various members of the subfamily of glutamate ionotropic receptors.  相似文献   

20.
The 90-kDa heat shock protein (Hsp90) is a molecular chaperone that assists both in ATP-independent sequestration of damaged proteins, and in ATP-dependent folding of numerous targets, such as nuclear hormone receptors and protein kinases. Recent work from our lab and others has established the existence of a second, C-terminal nucleotide binding site besides the well characterized N-terminal, geldanamycin-sensitive ATP-binding site. The cryptic C-terminal site becomes open only after the occupancy of the N-terminal site. Our present work demonstrates the applicability of the oxidative nucleotide affinity cleavage in the site-specific characterization of nucleotide binding proteins. We performed a systematic analysis of the nucleotide binding specificity of the Hsp90 nucleotide binding sites. N-terminal binding is specific to adenosine nucleotides with an intact adenine ring. Nicotinamide adenine dinucleotides and diadenosine polyphosphate alarmones are specific N-terminal nucleotides. The C-terminal binding site is much more unspecific-it interacts with both purine and pirimidine nucleotides. Efficient binding to the C-terminal site requires both charged residues and a larger hydrophobic moiety. GTP and UTP are specific C-terminal nucleotides. 2',3'-O-(2,4,6-trinitrophenyl)-nucleotides (TNP-ATP, TNP-GTP) and pyrophosphate access the C-terminal binding site without the need for an occupied N-terminal site. Our data provide additional evidence for the dynamic domain-domain interactions of Hsp90, give hints for the design of novel types of specific Hsp90 inhibitors, and raise the possibility that besides ATP, other small molecules might also interact with the C-terminal nucleotide binding site 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