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 390 毫秒
1.
We have cloned the M and S genes of the restriction-modification (R-M) system AhdI and have purified the resulting methyltransferase to homogeneity. M.AhdI is found to form a 170 kDa tetrameric enzyme having a subunit stoichiometry M2S2 (where the M and S subunits are responsible for methylation and DNA sequence specificity, respectively). Sedimentation equilibrium experiments show that the tetrameric enzyme dissociates to form a heterodimer at low concentration, with Kd ≈ 2 µM. The intact (tetrameric) enzyme binds specifically to a 30 bp DNA duplex containing the AhdI recognition sequence GACN5GTC with high affinity (Kd ≈ 50 nM), but at low enzyme concentration the DNA binding activity is governed by the dissociation of the tetramer into dimers, leading to a sigmoidal DNA binding curve. In contrast, only non-specific binding is observed if the duplex lacks the recognition sequence. Methylation activity of the purified enzyme was assessed by its ability to prevent restriction by the cognate endonuclease. The subunit structure of the M.AhdI methyltransferase resembles that of type I MTases, in contrast to the R.AhdI endonuclease which is typical of type II systems. AhdI appears to be a novel R-M system with properties intermediate between simple type II systems and more complex type I systems, and may represent an intermediate in the evolution of R-M systems.  相似文献   

2.
We have analyzed the structure and function of the integrin α1I domain harboring a gain-of-function mutation E317A. To promote protein crystallization, a double variant with an additional C139S mutation was used. In cell adhesion assays, the E317A mutation promoted binding to collagen. Similarly, the double mutation C139S/E317A increased adhesion compared with C139S alone. Furthermore, soluble α1I C139S/E317A was a higher avidity collagen binder than α1I C139S, indicating that the double variant represents an activated form. The crystal structure of the activated variant of α1I was solved at 1.9 Å resolution. The E317A mutation results in the unwinding of the αC helix, but the metal ion has moved toward loop 1, instead of loop 2 in the open α2I. Furthermore, unlike in the closed αI domains, the metal ion is pentacoordinated and, thus, prepared for ligand binding. Helix 7, which has moved downward in the open α2I structure, has not changed its position in the activated α1I variant. During the integrin activation, Glu335 on helix 7 binds to the metal ion at the metal ion-dependent adhesion site (MIDAS) of the β1 subunit. Interestingly, in our cell adhesion assays E317A could activate collagen binding even after mutating Glu335. This indicates that the stabilization of helix 7 into its downward position is not required if the α1 MIDAS is already open. To conclude, the activated α1I domain represents a novel conformation of the αI domain, mimicking the structural state where the Arg287-Glu317 ion pair has just broken during the integrin activation.  相似文献   

3.
Type-I DNA restriction–modification (R/M) systems are important agents in limiting the transmission of mobile genetic elements responsible for spreading bacterial resistance to antibiotics. EcoKI, a Type I R/M enzyme from Escherichia coli, acts by methylation- and sequence-specific recognition, leading to either methylation of DNA or translocation and cutting at a random site, often hundreds of base pairs away. Consisting of one specificity subunit, two modification subunits, and two DNA translocase/endonuclease subunits, EcoKI is inhibited by the T7 phage antirestriction protein ocr, a DNA mimic. We present a 3D density map generated by negative-stain electron microscopy and single particle analysis of the central core of the restriction complex, the M.EcoKI M2S1 methyltransferase, bound to ocr. We also present complete atomic models of M.EcoKI in complex with ocr and its cognate DNA giving a clear picture of the overall clamp-like operation of the enzyme. The model is consistent with a large body of experimental data on EcoKI published over 40 years.  相似文献   

4.
Over the past few years, it has become clear that an important mechanism by which large-conductance Ca2+-activated K+ channel (BKCa) activity is regulated is the tissue-specific expression of auxiliary β subunits. The first of these to be identified, β1, is expressed predominately in smooth muscle and causes dramatic effects, increasing the apparent affinity of the channel for Ca2+ 10-fold at 0 mV, and shifting the range of voltages over which the channel activates −80 mV at 9.1 μM Ca2+. With this study, we address the question: which aspects of BKCa gating are altered by β1 to bring about these effects: Ca2+ binding, voltage sensing, or the intrinsic energetics of channel opening? The approach we have taken is to express the β1 subunit together with the BKCa α subunit in Xenopus oocytes, and then to compare β1''s steady state effects over a wide range of Ca2+ concentrations and membrane voltages to those predicted by allosteric models whose parameters have been altered to mimic changes in the aspects of gating listed above. The results of our analysis suggest that much of β1''s steady state effects can be accounted for by a reduction in the intrinsic energy the channel must overcome to open and a decrease in its voltage sensitivity, with little change in the affinity of the channel for Ca2+ when it is either open or closed. Interestingly, however, the small changes in Ca2+ binding affinity suggested by our analysis (Kc 7.4 μM → 9.6 μM; Ko = 0.80 μM → 0.65 μM) do appear to be functionally important. We also show that β1 affects the mSlo conductance–voltage relation in the essential absence of Ca2+, shifting it +20 mV and reducing its apparent gating charge 38%, and we develop methods for distinguishing between alterations in Ca2+ binding and other aspects of BKCa channel gating that may be of general use.  相似文献   

5.
Movements of native C505 during channel gating in CNGA1 channels   总被引:1,自引:1,他引:0  
We investigated conformational changes occurring in the C-linker and cyclic nucleotide-binding (CNB) domain of CNGA1 channels by analyzing the inhibition induced by thiol-specific reagents in mutant channels Q409C and A414C in the open and closed state. Cd2+ (200 μM) inhibited irreversibly mutant channels Q409C and A414C in the closed but not in the open state. Cd2+ inhibition was abolished in the mutant A414Ccys-free, in the double mutant A414C + C505T and in the tandem construct A414C + C505T/CNGA1, but it was present in the construct A414C + C505cys-free. The cross-linker reagent M-2-M inhibited mutant channel Q409C in the open state. M-2-M inhibition in the open state was abolished in the double mutant Q409C + C505T and in the tandem construct Q409C + C505T/CNGA1. These results show that Cα of C505 in the closed state is located at a distance between 4 and 10.5 Å from the Cα of A414 of the same subunit, but in the open state C505 moves towards Q409 of the same subunit at a distance that ranges from 10.5 to 12.3 Å from Cα of this residue. These results are not consistent with a 3-D structure of the CNGA1 channel homologous to the structure of HCN2 channels either in the open or in the closed state.  相似文献   

6.
P2X receptors are ligand-gated cation channels that transition from closed to open states upon binding ATP. The crystal structure of the closed zebrafish P2X4.1 receptor directly reveals that the ion-conducting pathway is formed by three transmembrane domain 2 (TM2) α-helices, each being provided by the three subunits of the trimer. However, the transitions in TM2 that accompany channel opening are incompletely understood and remain unresolved. In this study, we quantified gated access to Cd2+ at substituted cysteines in TM2 of P2X2 receptors in the open and closed states. Our data for the closed state are consistent with the zebrafish P2X4.1 structure, with isoleucines and threonines (Ile-332 and Thr-336) positioned one helical turn apart lining the channel wall on approach to the gate. Our data for the open state reveal gated access to deeper parts of the pore (Thr-339, Val-343, Asp-349, and Leu-353), suggesting the closed channel gate is between Thr-336 and Thr-339. We also found unexpected interactions between native Cys-348 and D349C that result in tight Cd2+ binding deep within the intracellular vestibule in the open state. Interpreted with a P2X2 receptor structural model of the closed state, our data suggest that the channel gate opens near Thr-336/Thr-339 and is accompanied by movement of the pore-lining regions, which narrow toward the cytosolic end of TM2 in the open state. Such transitions would relieve the barrier to ion flow and render the intracellular vestibule less splayed during channel opening in the presence of ATP.  相似文献   

7.
8.
9.
We have determined the structure of Pvu II methyltransferase (M. Pvu II) complexed with S -adenosyl-L-methionine (AdoMet) by multiwavelength anomalous diffraction, using a crystal of the selenomethionine-substituted protein. M. Pvu II catalyzes transfer of the methyl group from AdoMet to the exocyclic amino (N4) nitrogen of the central cytosine in its recognition sequence 5'-CAGCTG-3'. The protein is dominated by an open alpha/beta-sheet structure with a prominent V-shaped cleft: AdoMet and catalytic amino acids are located at the bottom of this cleft. The size and the basic nature of the cleft are consistent with duplex DNA binding. The target (methylatable) cytosine, if flipped out of the double helical DNA as seen for DNA methyltransferases that generate 5-methylcytosine, would fit into the concave active site next to the AdoMet. This M. Pvu IIalpha/beta-sheet structure is very similar to those of M. Hha I (a cytosine C5 methyltransferase) and M. Taq I (an adenine N6 methyltransferase), consistent with a model predicting that DNA methyltransferases share a common structural fold while having the major functional regions permuted into three distinct linear orders. The main feature of the common fold is a seven-stranded beta-sheet (6 7 5 4 1 2 3) formed by five parallel beta-strands and an antiparallel beta-hairpin. The beta-sheet is flanked by six parallel alpha-helices, three on each side. The AdoMet binding site is located at the C-terminal ends of strands beta1 and beta2 and the active site is at the C-terminal ends of strands beta4 and beta5 and the N-terminal end of strand beta7. The AdoMet-protein interactions are almost identical among M. Pvu II, M. Hha I and M. Taq I, as well as in an RNA methyltransferase and at least one small molecule methyltransferase. The structural similarity among the active sites of M. Pvu II, M. Taq I and M. Hha I reveals that catalytic amino acids essential for cytosine N4 and adenine N6 methylation coincide spatially with those for cytosine C5 methylation, suggesting a mechanism for amino methylation.  相似文献   

10.
Enzymes that modify DNA are faced with significant challenges in specificity for both substrate binding and catalysis. We describe how single hydrogen bonds between M.HhaI, a DNA cytosine methyltransferase, and its DNA substrate regulate the positioning of a peptide loop which is ∼28 Å away. Stopped-flow fluorescence measurements of a tryptophan inserted into the loop provide real-time observations of conformational rearrangements. These long-range interactions that correlate with substrate binding and critically, enzyme turnover, will have broad application to enzyme specificity and drug design for this medically relevant class of enzymes.Sequence-specific modification of DNA is essential for nearly all forms of life and contributes to a myriad of biological processes including gene regulation, mismatch repair, host defense, DNA replication, and genetic imprinting. Methylation of cytosine and adenine bases is a key epigenetic process whereby phenotypic changes are inherited without altering the DNA sequence (1). The central role of the bacterial and mammalian S-adenosylmethionine (AdoMet)2-dependent DNA methyltransferases in virulence regulation and tumorigenesis, respectively, have led these enzymes to be validated targets for antibiotic and cancer therapies (2, 3). However, AdoMet-dependent enzymes catalyze diverse reactions, and the design of potent and selective DNA methyltransferase inhibitors is particularly challenging (4, 5). The design of drugs that bind outside the active site is a particularly attractive means of inhibition for enzymes with common cofactors like AdoMet because off-target inhibition often leads to toxicity (6). Unfortunately, robust methods to identify and characterize such critical binding sites distal from the active site have not been developed.DNA methyltransferases bind to a particular DNA sequence, stabilize the target base into an extrahelical position within the enzyme active site, and transfer the methyl moiety from AdoMet to the DNA (7). During this process, dramatic changes in the DNA structure such as bending, base flipping, or the intercalation of residues into the recognition sequence are often accompanied by large scale protein rearrangements (8). Here we characterized a specific conformational rearrangement of M.HhaI, a model DNA cytosine C5 methyltransferase with a cognate recognition sequence of 5′-GCGC-3′. Many structures of M.HhaI are available at high resolution including an ensemble of complexes with either cognate or nonspecific DNA (9, 10). Reorganization of an essential catalytic loop (residues 80–100) is regulated by sequence-specific protein-DNA interactions that occur ∼28 Å away from the catalytic loop (Fig. 1). Our work quantifies the importance of such distal communication in sequence-specific DNA modification and provides plausible structural mechanisms.Open in a separate windowFIGURE 1.Loop interactions in M.HhaI. A, two superimposed structures of M.HhaI are shown with the catalytic loop highlighted. Enzymes are in blue (open conformer) and red (closed conformer) with the cofactor in orange, the flipped cytosine in green, and position 1 of the DNA recognition sequence colored by atom along with Cys-81, Ile-86, and Arg-240. The large blue arrow shows the long-range structural communication between Arg-240 and the catalytic loop. B, close-up of the interactions between Arg-240, Ile-86, and position 1 of the recognition sequence. The flipped cytosine is in green. Removal of N2 from the guanosine with an inosine base maintains near cognate loop motion while removal of O6 with a 2AP base has almost no loop motion.DNA-dependent positioning of the catalytic loop in M.HhaI was first observed crystallographically; cognate DNA stabilizes the loop-closed conformer, while nonspecific DNA leaves the loop in the open conformer (9, 10). Correct positioning of this loop is essential for catalysis because C81, the active site nucleophile that attacks the target cytosine base at the C6 position (supplemental Fig. S1), is ∼9.6 Å away in the loop-open conformer (Fig. 1A). Populating the closed conformer of the loop is essential for tight DNA binding and stabilizing the target cytosine that is flipped out of the DNA duplex (1113). Using stopped-flow fluorescence spectroscopy to monitor the environment of tryptophan (Trp) residues inserted into the catalytic loop, we recently observed reorganization of this loop upon DNA binding in the absence of cofactor using the M.HhaI mutants W41F, W41F/K91W, and W41F/E94W (12). Loop positioning and the interconversion between the open and closed conformers, as determined from the intensities and rates of change in fluorescence signal are highly dependent on DNA sequence and confirm that cognate DNA stabilizes the loop-closed conformer whereas nonspecific DNA stabilizes the open conformer.In this study, W41F/K91W and W41F/E94W M.HhaI were preincubated with cognate (COG), non-cognate (NC), or nonspecific (NS) DNA and mixed with cofactor or cofactor product, AdoMet and S-adenosylhomocysteine (AdoHcy), respectively, in a stopped-flow apparatus. Differences in observed fluorescence intensity are indicative of shifts in the populations of the various loop conformers; no observable fluorescence change suggests no significant change in population and thus, essentially no loop positioning to the closed conformer. Non-cognate and cognate sequences are nearly identical but differ by a single base change, whereas the nonspecific DNA substrate has no similarity to the cognate sequence. As a methyltransferase searches for the cognate site within a genome, it must encounter both non-cognate and nonspecific DNA sequences and be able to distinguish these from the cognate sequence. We examine both binding, using the cofactor product AdoHcy, and catalysis, using the AdoMet cofactor of M.HhaI, with these various DNA substrates.  相似文献   

11.
I A Taylor  K G Davis  D Watts    G G Kneale 《The EMBO journal》1994,13(23):5772-5778
The type IC DNA methyltransferase M.EcoR124I is a complex multisubunit enzyme that recognizes the non-palindromic DNA sequence GAAN6RTCG. Small angle X-ray scattering has been used to investigate the solution structure of the methyltransferase and of complexes of the enzyme with unmethylated and hemimethylated 30 bp DNA duplexes containing the specific recognition sequence. A major change in the quaternary structure of the enzyme is observed following DNA binding, based on a decrease in the radius of gyration from 56 to 40 A and a reduction in the maximum dimension of the enzyme from 180 to 112 A. The structural transition observed is independent of the methylation state of the DNA. CD shows that there is no change in the secondary structure of the protein subunits when DNA is bound. In contrast, there is a large increase in the CD signal arising from the DNA, suggesting considerable structural distortion which may allow access to the bases targeted for methylation. We propose that DNA binding induces a large rotation of the two HsdM subunits towards the DNA, mediated by hinge bending domains in the specificity subunit HsdS.  相似文献   

12.
Type I restriction-modification (R-M) enzymes are composed of three different subunits, of which HsdS determines DNA specificity, HsdM is responsible for DNA methylation and HsdR is required for restriction. The HsdM and HsdS subunits can also form an independent DNA methyltransferase with a subunit stoichiometry of M2S1. We found that the purified Eco R124I R-M enzyme was a mixture of two species as detected by the presence of two differently migrating specific DNA-protein complexes in a gel retardation assay. An analysis of protein subunits isolated from the complexes indicated that the larger species had a stoichiometry of R2M2S1and the smaller species had a stoichiometry of R1M2S1. In vitro analysis of subunit assembly revealed that while binding of the first HsdR subunit to the M2S1complex was very tight, the second HsdR subunit was bound weakly and it dissociated from the R1M2S1complex with an apparent K d of approximately 2.4 x 10(-7) M. Functional assays have shown that only the R2M2S1complex is capable of DNA cleavage, however, the R1M2S1complex retains ATPase activity. The relevance of this situation is discussed in terms of the regulation of restriction activity in vivo upon conjugative transfer of a plasmid-born R-M system into an unmodified host cell.  相似文献   

13.
RNase H2 cleaves RNA sequences that are part of RNA/DNA hybrids or that are incorporated into DNA, thus, preventing genomic instability and the accumulation of aberrant nucleic acid, which in humans induces Aicardi-Goutières syndrome, a severe autoimmune disorder. The 3.1 Å crystal structure of human RNase H2 presented here allowed us to map the positions of all 29 mutations found in Aicardi-Goutières syndrome patients, several of which were not visible in the previously reported mouse RNase H2. We propose the possible effects of these mutations on the protein stability and function. Bacterial and eukaryotic RNases H2 differ in composition and substrate specificity. Bacterial RNases H2 are monomeric proteins and homologs of the eukaryotic RNases H2 catalytic subunit, which in addition possesses two accessory proteins. The eukaryotic RNase H2 heterotrimeric complex recognizes RNA/DNA hybrids and (5′)RNA-DNA(3′)/DNA junction hybrids as substrates with similar efficiency, whereas bacterial RNases H2 are highly specialized in the recognition of the (5′)RNA-DNA(3′) junction and very poorly cleave RNA/DNA hybrids in the presence of Mg2+ ions. Using the crystal structure of the Thermotoga maritima RNase H2-substrate complex, we modeled the human RNase H2-substrate complex and verified the model by mutational analysis. Our model indicates that the difference in substrate preference stems from the different position of the crucial tyrosine residue involved in substrate binding and recognition.  相似文献   

14.
Glycine N-methyltransferase (S-adenosyl-l-methionine: glycine methyltransferase, EC 2.1.1.20; GNMT) catalyzes the AdoMet-dependent methylation of glycine to form sarcosine (N-methylglycine). Unlike most methyltransferases, GNMT is a tetrameric protein showing a positive cooperativity in AdoMet binding and weak inhibition by S-adenosylhomocysteine (AdoHcy). The first crystal structure of GNMT complexed with AdoMet showed a unique "closed" molecular basket structure, in which the N-terminal section penetrates and corks the entrance of the adjacent subunit. Thus, the apparent entrance or exit of the active site is not recognizable in the subunit structure, suggesting that the enzyme must possess a second, enzymatically active, "open" structural conformation. A new crystalline form of the R175K enzyme has been grown in the presence of an excess of AdoHcy, and its crystal structure has been determined at 3.0 A resolution. In this structure, the N-terminal domain (40 amino acid residues) of each subunit has moved out of the active site of the adjacent subunit, and the entrances of the active sites are now opened widely. An AdoHcy molecule has entered the site occupied in the "closed" structure by Glu15 and Gly16 of the N-terminal domain of the adjacent subunit. An AdoHcy binds to the consensus AdoMet binding site observed in the other methyltransferase. This AdoHcy binding site supports the glycine binding site (Arg175) deduced from a chemical modification study and site-directed mutagenesis (R175K). The crystal structures of WT and R175K enzymes were also determined at 2.5 A resolution. These enzyme structures have a closed molecular basket structure and are isomorphous to the previously determined AdoMet-GNMT structure. By comparing the open structure to the closed structure, mechanisms for auto-inhibition and for the forced release of the product AdoHcy have been revealed in the GNMT structure. The N-terminal section of the adjacent subunit occupies the AdoMet binding site and thus inhibits the methyltransfer reaction, whereas the same N-terminal section forces the departure of the potentially potent inhibitor AdoHcy from the active site and thus facilitates the methyltransfer reaction. Consequently GNMT is less active at a low level of AdoMet concentration, and is only weakly inhibited by AdoHcy. These properties of GNMT are particularly suited for regulation of the cellular AdoMet/AdoHcy ratio.  相似文献   

15.
Photoaffinity labeling of γ-aminobutyric acid type A (GABAA)-receptors (GABAAR) with an etomidate analog and mutational analyses of direct activation of GABAAR by neurosteroids have each led to the proposal that these structurally distinct general anesthetics bind to sites in GABAARs in the transmembrane domain at the interface between the β and α subunits. We tested whether the two ligand binding sites might overlap by examining whether neuroactive steroids inhibited etomidate analog photolabeling. We previously identified (Li, G. D., Chiara, D. C., Sawyer, G. W., Husain, S. S., Olsen, R. W., and Cohen, J. B. (2006) J. Neurosci. 26, 11599–11605) azietomidate photolabeling of GABAAR α1Met-236 and βMet-286 (in αM1 and βM3). Positioning these two photolabeled amino acids in a single type of binding site at the interface of β and α subunits (two copies per pentamer) is consistent with a GABAAR homology model based upon the structure of the nicotinic acetylcholine receptor and with recent αM1 to βM3 cross-linking data. Biologically active neurosteroids enhance rather than inhibit azietomidate photolabeling, as assayed at the level of GABAAR subunits on analytical SDS-PAGE, and protein microsequencing establishes that the GABAAR-modulating neurosteroids do not inhibit photolabeling of GABAAR α1Met-236 or βMet-286 but enhance labeling of α1Met-236. Thus modulatory steroids do not bind at the same site as etomidate, and neither of the amino acids identified as neurosteroid activation determinants (Hosie, A. M., Wilkins, M. E., da Silva, H. M., and Smart, T. G. (2006) Nature 444, 486–489) are located at the subunit interface defined by our etomidate site model.GABAA3 receptors (GABAAR) are major mediators of brain inhibitory neurotransmission and participate in most circuits and behavioral pathways relevant to normal and pathological function (1). GABAAR are subject to modulation by endogenous neurosteroids, as well as myriad clinically important central nervous system drugs including general anesthetics, benzodiazepines, and possibly ethanol (1, 2). The mechanism of GABAAR modulation by these different classes of drugs is of major interest, including identification of the receptor amino acid residues involved in binding and action of the drugs.In the absence of high resolution crystal structures of drug-receptor complexes, the locations of anesthetic binding sites in GABAARs have been predicted based upon analyses of functional properties of point mutant receptors, which identified residues in the α and β subunit M1–M4 transmembrane helices important for modulation by volatile anesthetics (primarily α subunit) and by intravenous agents, including etomidate and propofol (β subunit) (35). Position βM2–15, numbered relative to the N terminus of the helix, functions as a major determinant of etomidate and propofol potency as GABA modulators in vitro and in vivo (68). By contrast, this residue is not implicated for modulation by the neurosteroids, potent endogenous modulators of GABAAR (9).Photoaffinity labeling, which allows the identification of residues in proximity to drug binding sites (10, 11), has been used to identify two GABAAR amino acids covalently modified by the etomidate analog [3H]azietomidate (12): α1Met-236 within αM1 and βMet-286 within βM3. Photolabeling of these residues was inhibited equally by nonradioactive etomidate and enhanced proportionately by GABA present in the assay, consistent with the presence of these two residues in a common drug binding pocket that would be located at the interface between the β and α subunits in the transmembrane domain (12). Mutational analyses identify these positions as etomidate and propofol sensitivity determinants (1315).A recent mutagenesis study (16) identified two other residues in GABAAR αM1 and βM3 as critical for direct activation by neurosteroids, αThr-236 (rat numbering, corresponding to α1Thr-237, bovine numbering used here and by Li et al. (12))4 and βTyr-284. These residues were also proposed to contribute to a neurosteroid binding pocket in the transmembrane domain at the interface between β and α subunits, based upon their location in an alternative GABAAR structural model that positioned those amino acids, and not α1Met-236 or βMet-286, at the subunit interface. For GABAARs and other members of the Cys-loop superfamily of neurotransmitter-gated ion channels, the transmembrane domain of each subunit is made up of a loose bundle of four α helices (M1–M4), with M2 from each subunit contributing to the lumen of the ion channel and M4 positioned peripherally in greatest contact with lipid, as seen in the structures of the Torpedo nicotinic acetylcholine receptor (nAChR) (17) and in distantly related prokaryote homologs (18). However, uncertainties in the alignment of GABAAR subunit sequences relative to those of the nAChR result in alternative GABAAR homology models (12, 19, 20) that differ in the location of amino acids in the M3 and M4 membrane-spanning helices and in the M1 helix in some models (16, 21).If etomidate and neurosteroids both bind at the same β/α interface in the GABAAR transmembrane domain, the limited space available for ligand binding suggests that their binding pockets might overlap and that ligand binding would be mutually exclusive. To address this question, we photolabeled purified bovine brain GABAAR with [3H]azietomidate in the presence of different neuroactive steroids and determined by protein microsequencing whether active neurosteroids inhibited labeling of α1Met-236 and βMet-286, as expected for mutually exclusive binding, or resulted in [3H]azietomidate photolabeling of other amino acids, a possible consequence of allosteric interactions. Active steroids failed to inhibit labeling and enhanced labeling of α1Met-236, clearly indicating that the neurosteroid and the etomidate sites are distinct. Our GABAAR homology model that positions α1Met-236 and βMet-286 at the β/α interface, but not that of Hosie et al. (16), is also consistent with cysteine substitution cross-linking studies (20, 22), which define the proximity relations between amino acids in the αM1, αM2, αM3, and βM3 helices, and these results support the interpretation that the two residues photolabeled by [3H]azietomidate are part of a single subunit interface binding pocket, whereas the steroid sensitivity determinants identified by mutagenesis neither are at the β/α subunit interface nor are contributors to a common binding pocket.  相似文献   

16.
The J-binding protein 1 (JBP1) is essential for biosynthesis and maintenance of DNA base-J (β-d-glucosyl-hydroxymethyluracil). Base-J and JBP1 are confined to some pathogenic protozoa and are absent from higher eukaryotes, prokaryotes and viruses. We show that JBP1 recognizes J-containing DNA (J-DNA) through a 160-residue domain, DB-JBP1, with 10 000-fold preference over normal DNA. The crystal structure of DB-JBP1 revealed a helix-turn-helix variant fold, a ‘helical bouquet’ with a ‘ribbon’ helix encompassing the amino acids responsible for DNA binding. Mutation of a single residue (Asp525) in the ribbon helix abrogates specificity toward J-DNA. The same mutation renders JBP1 unable to rescue the targeted deletion of endogenous JBP1 genes in Leishmania and changes its distribution in the nucleus. Based on mutational analysis and hydrogen/deuterium-exchange mass-spectrometry data, a model of JBP1 bound to J-DNA was constructed and validated by small-angle X-ray scattering data. Our results open new possibilities for targeted prevention of J-DNA recognition as a therapeutic intervention for parasitic diseases.  相似文献   

17.
18.
The EcoKI DNA methyltransferase is a trimeric protein comprised of two modification subunits (M) and one sequence specificity subunit (S). This enzyme forms the core of the EcoKI restriction/modification (RM) enzyme. The 3′ end of the gene encoding the M subunit overlaps by 1 nt the start of the gene for the S subunit. Translation from the two different open reading frames is translationally coupled. Mutagenesis to remove the frameshift and fuse the two subunits together produces a functional RM enzyme in vivo with the same properties as the natural EcoKI system. The fusion protein can be purified and forms an active restriction enzyme upon addition of restriction subunits and of additional M subunit. The Type I RM systems are grouped into families, IA to IE, defined by complementation, hybridization and sequence similarity. The fusion protein forms an evolutionary intermediate form lying between the Type IA family of RM enzymes and the Type IB family of RM enzymes which have the frameshift located at a different part of the gene sequence.  相似文献   

19.

Background

Methylenetetrahydrofolate reductase (MTHFR) is one of the enzymes involved in homocysteine metabolism. Despite considerable genetic and clinical attention, the reaction mechanism and regulation of this enzyme are not fully understood because of difficult production and poor stability. While recombinant enzymes from thermophilic organisms are often stable and easy to prepare, properties of thermostable MTHFRs have not yet been reported.

Methodology/Principal Findings

MTHFR from Thermus thermophilus HB8, a homologue of Escherichia coli MetF, has been expressed in E. coli and purified. The purified MTHFR was chiefly obtained as a heterodimer of apo- and holo-subunits, that is, one flavin adenine dinucleotide (FAD) prosthetic group bound per dimer. The crystal structure of the holo-subunit was quite similar to the β8α8 barrel of E. coli MTHFR, while that of the apo-subunit was a previously unobserved closed form. In addition, the intersubunit interface of the dimer in the crystals was different from any of the subunit interfaces of the tetramer of E. coli MTHFR. Free FAD could be incorporated into the apo-subunit of the purified Thermus enzyme after purification, forming a homodimer of holo-subunits. Comparison of the crystal structures of the heterodimer and the homodimer revealed different intersubunit interfaces, indicating a large conformational change upon FAD binding. Most of the biochemical properties of the heterodimer and the homodimer were the same, except that the homodimer showed ≈50% activity per FAD-bound subunit in folate-dependent reactions.

Conclusions/Significance

The different intersubunit interfaces and rearrangement of subunits of Thermus MTHFR may be related to human enzyme properties, such as the allosteric regulation by S-adenosylmethionine and the enhanced instability of the Ala222Val mutant upon loss of FAD. Whereas E. coli MTHFR was the only structural model for human MTHFR to date, our findings suggest that Thermus MTHFR will be another useful model for this important enzyme.  相似文献   

20.
The 5′-cap of spliceosomal small nuclear RNAs, some small nucleolar RNAs and of telomerase RNA was found to be hypermethylated in vivo. The Trimethylguanosine Synthase 1 (TGS1) mediates this conversion of the 7-methylguanosine-cap to the 2,2,7-trimethylguanosine (m3G)-cap during maturation of the RNPs. For mammalian UsnRNAs the generated m2,2,7G-cap is one part of a bipartite import signal mediating the transport of the UsnRNP-core complex into the nucleus. In order to understand the structural organization of human TGS1 as well as substrate binding and recognition we solved the crystal structure of the active TGS1 methyltransferase domain containing both, the minimal substrate m7GTP and the reaction product S-adenosyl-l-homocysteine (AdoHcy). The methyltransferase of human TGS1 harbors the canonical class 1 methyltransferase fold as well as an unique N-terminal, α-helical domain of 40 amino acids, which is essential for m7G-cap binding and catalysis. The crystal structure of the substrate bound methyltransferase domain as well as mutagenesis studies provide insight into the catalytic mechanism of TGS1.  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