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 15 毫秒
1.
Proteorhodopsin is an ion-translocating member of the microbial rhodopsin family. Light absorption by its retinal chromophore initiates a photocycle, driven by trans/cis isomerization, leading to transmembrane translocation of a proton toward the extracellular side of the cytoplasmic membrane. Here we report a study on the photoisomerization dynamics of the retinal chromophore of proteorhodopsin, using femtosecond time-resolved spectroscopy, by probing in the visible- and in the midinfrared spectral regions. Experiments were performed both at pH 9.5 (a physiologically relevant pH value in which the primary proton acceptor of the protonated Schiff base, Asp97, is deprotonated) and at pH 6.5 (with Asp97 protonated). Simultaneous analysis of the data sets recorded in the two spectral regions and at both pH values reveals a multiexponential excited state decay, with time constants of ∼0.2 ps, ∼2 ps, and ∼20 ps. From the difference spectra associated with these dynamics, we conclude that there are two chromophore-isomerizaton pathways that lead to the K-state: one with an effective rate of ∼(2 ps)−1 and the other with a rate of ∼(20 ps)−1. At high pH, both pathways are equally effective, with an estimated quantum yield for K-formation of ∼0.7. At pH 6.5, the slower pathway is less productive, which results in an isomerization quantum yield of 0.5. We further observe an ultrafast response of residue Asp227, which forms part of the counterion complex, corresponding to a strengthening of its hydrogen bond with the Schiff base on K-state formation; and a feature that develops on the 0.2 ps and 2 ps timescale and probably reflects a response of an amide II band in reaction to the isomerization process.  相似文献   

2.
The genome of thylakoidless cyanobacterium Gloeobacter violaceus encodes a fast-cycling rhodopsin capable of light-driven proton transport. We characterize the dark state, the photocycle, and the proton translocation pathway of GR spectroscopically. The dark state of GR contains predominantly all-trans-retinal and, similar to proteorhodopsin, does not show the light/dark adaptation. We found an unusually strong coupling between the conformation of the retinal and the site of Glu132, the homolog of Asp96 of BR. Although the photocycle of GR is similar to that of proteorhodopsin in general, it differs in accumulating two intermediates typical for BR, the L-like and the N-like states. The latter state has a deprotonated cytoplasmic proton donor and is spectrally distinct from the strongly red-shifted N intermediate known for proteorhodopsin. The proton uptake precedes the release and occurs during the transition to the O intermediate. The proton translocation pathway of GR is similar to those of other proton-pumping rhodopsins, involving homologs of BR Schiff base proton acceptor and donor Asp85 and Asp96 (Asp121 and Glu132). We assigned a pair of FTIR bands (positive at 1749 cm−1 and negative at 1734 cm−1) to the protonation and deprotonation, respectively, of these carboxylic acids.  相似文献   

3.
Ranga Partha 《BBA》2005,1708(1):6-12
In wild-type proteorhodopsin (pR), titration of the chromophore's counterion Asp97 occurs with a pKa of 8.2±0.1. R94C mutation reduces this slightly to 7.0±0.2, irrespective of treatment with ethylguanidinium. This contrasts with the homologous archaeal protein bacteriorhodopsin (bR), where R82C mutation was previously shown to elevate the pKa of Asp85 by ∼5 units, while reconstitution with ethylguanidinium restores it nearly to the wild-type value of 2.5. We conclude there is much weaker electrostatic coupling between Arg94 and Asp97 in the unphotolyzed state of pR, in comparison to Arg82 and Asp85 in bR. Therefore, while fast light-driven H+ release may depend on these two residues in pR as in bR, no tightly conserved pre-photolysis configuration of them is required.  相似文献   

4.
HD-domain phosphohydrolases have nucleotidase and phosphodiesterase activities and play important roles in the metabolism of nucleotides and in signaling. We present three 2.1-Å-resolution crystal structures (one in the free state and two complexed with natural substrates) of an HD-domain phosphohydrolase, the Escherichia coli 5′-nucleotidase YfbR. The free-state structure of YfbR contains a large cavity accommodating the metal-coordinating HD motif (H33, H68, D69, and D137) and other conserved residues (R18, E72, and D77). Alanine scanning mutagenesis confirms that these residues are important for activity. Two structures of the catalytically inactive mutant E72A complexed with Co2+ and either thymidine-5′-monophosphate or 2′-deoxyriboadenosine-5′-monophosphate disclose the novel binding mode of deoxyribonucleotides in the active site. Residue R18 stabilizes the phosphate on the Co2+, and residue D77 forms a strong hydrogen bond critical for binding the ribose. The indole side chain of W19 is located close to the 2′-carbon atom of the deoxyribose moiety and is proposed to act as the selectivity switch for deoxyribonucleotide, which is supported by comparison to YfdR, another 5′-nucleotidase in E. coli. The nucleotide bases of both deoxyriboadenosine-5′-monophosphate and thymidine-5′-monophosphate make no specific hydrogen bonds with the protein, explaining the lack of nucleotide base selectivity. The YfbR E72A substrate complex structures also suggest a plausible single-step nucleophilic substitution mechanism. This is the first proposed molecular mechanism for an HD-domain phosphohydrolase based directly on substrate-bound crystal structures.  相似文献   

5.
Five new octahedral iron(II) complexes [FeL2(4-dpa)]n(EtOH) (1), [FeL2(bipy)]n(DMF) (2), [FeL1(bpee)]n (3), [Fe2L3(1-meim)4](1-meim)4 (4) and [FeL1(DMAP)2] (5), with L1 and L2 being tetradentate coordinating Schiff base like ligands (L1 = (E,E)-[{diethyl-2,2′-[1,2-phenylenebis(iminomethylidyne)]bis[3-oxobutanato](2-)-N,N′,O3,O3′}, L2 = (3,3′)-[{1,2-phenylenebis(iminomethylidyne)]bis(2,4-pentane-dionato)(2-)-N,N′,O2,O2′}) and L3 being a octadentate dinucleating coordinating Schiff base like ligand ({tetraethyl-(E,E,E,E)-2,2′,2′′,2′′′-[1,2,4,5-phenylentetra(iminomethylidine)]tetra[3-oxobutanoato](2-)-N,N′,N′′,N′′′,O3,O3′,O3′′,O3′′′}); 4-dpa = di(4-picolyl)-amine, bipy = 4,4′-bipyridine, bpee = trans-1,2-bis(4-pyridyl)ethylene, 1-meim = 1-methylimidazole and DMAP = 4-dimethylaminopyridine, have been synthesized and characterised using X-ray structure analysis and T-dependent susceptibility measurements. Both methods indicate that all iron(II) centres are in the paramagnetic high-spin state over the whole temperature range investigated. The O-Fe-O angle, the so called bit of the equatorial ligand, is with an average of 111° in the region typical for high-spin iron(II) complexes of this ligand type. In the case of compound 1 an infinite two-dimensional hydrogen bond network can be found, for the compounds 2-4 no hydrogen bond interactions are observed between the complex molecules. A comparison of the curve progression obtained from the magnetic measurements of the mononuclear complex 5 and the polymeric complexes 1-3 leads to the conclusion that no magnetic interactions are mediated over the bridging axial ligands. For the dinuclear complex 4 weak antiferromagnetic interactions between the two iron centres are found.  相似文献   

6.
A subclass of proteins with the SEA (sea urchin sperm protein, enterokinase, and agrin) domain fold exists as heterodimers generated by autoproteolytic cleavage within a characteristic G− 1S+ 1VVV sequence. Autoproteolysis occurs by a nucleophilic attack of the serine hydroxyl on the vicinal glycine carbonyl followed by an N → O acyl shift and hydrolysis of the resulting ester. The reaction has been suggested to be accelerated by the straining of the scissile peptide bond upon protein folding. In an accompanying article, we report the mechanism; in this article, we provide further key evidence and account for the energetics of coupled protein folding and autoproteolysis. Cleavage of the GPR116 domain and that of the MUC1 SEA domain occur with half-life (t½) values of 12 and 18 min, respectively, with lowering of the free energy of the activation barrier by ∼ 10 kcal mol− 1 compared with uncatalyzed hydrolysis. The free energies of unfolding of the GPR116 and MUC1 SEA domains were measured to ∼ 11 and ∼ 15 kcal mol− 1, respectively, but ∼ 7 kcal mol− 1 of conformational energy is partitioned as strain over the scissile peptide bond in the precursor to catalyze autoproteolysis by substrate destabilization. A straining energy of ∼ 7 kcal mol− 1 was measured by using both a pre-equilibrium model to analyze stability and cleavage kinetics data obtained with the GPR116 SEA domain destabilized by core mutations or urea addition, as well as the difference in thermodynamic stabilities of the MUC1 SEA precursor mutant S1098A (with a G− 1A+ 1VVV motif) and the wild-type protein. The results imply that cleavage by N → O acyl shift alone would proceed with a t½ of ∼ 2.3 years, which is too slow to be biochemically effective. A subsequent review of structural data on other self-cleaving proteins suggests that conformational strain of the scissile peptide bond may be a common mechanism of autoproteolysis.  相似文献   

7.
The Schiff base ligands, N,N′-bis(2-hydroxyacetophenone)-1,2-diaminoethane (acphenH2) and N,N′-bis(2-hydroxyacetophenone)-1,3-diaminopropane (acphpnH2), prepared in situ were used to synthesise two new Mn(III) complexes which were characterised by crystallography and variable temperature magnetic measurements. [Mn(acphen)NCS]2 is a phenoxo-bridged dimeric compound with the thiocyanate coordinating in the usual bent mode (Mn-N-C angle, 152°) and is weakly antiferromagnetic. Since there are no significant inter-dimer contacts in the crystal, the low temperature magnetic behaviour is influenced by single ion zero-field splitting. Exact diagonalisation of the spin Hamiltonian was performed to derive the following parameters: J = −0.7 cm−1, D = −0.6 cm−1. Mn(acphpn)(H2O)NCS is monomeric with an unusual linearly coordinated thiocyanate (Mn-N-C angle, 178°). Two lattice water molecules link the Mn(III) complex molecules through hydrogen bonds to form one-dimensional chains in the crystal. Magnetic exchange along the chain makes this compound also weakly antiferromagnetic with J ∼ -2 cm−1.  相似文献   

8.
Cryptonemia seminervis biosynthesizes a family of d,l-hybrid galactans based on the classical 3-linked β-d-galactopyranosyl→4-linked α-d- and α-l-galactopyranosyl alternating sequence (A-units→B-units) with major amounts of α-d- and α-l-galactose and 3,6-anhydro-d- and l-galactose and lesser percentages of 3,6-anhydro-2-O-methyl-l-galactose, 2-O-methyl-, 4-O-methyl- and 6-O-methylgalactoses. The dispersion of structures in this family is based on five structural factors, namely: (a) the amount and position of substituent groups as sulfate (major), pyruvic acid ketals, methoxyl and glycosyl side-chain (4-O-methyl galactopyranosyl and/or xylosyl); (b) the ratio galactose/3,6-anhydrogalactose in the B-units; (c) the ratio d,l-galactoses and d,l-3,6-anhydrogalactoses also in the B-units, (d) the formation of diads and (e) the sequence of the diads in the linear backbone. Considering these variables it is not unexpected to find in the fractions studied at least 18 structural units producing highly complex structures. Structural studies carried out in two major fractions (S2S-3 and S2S-4) showed that these galactans were formed mainly by β-d-galactopyranosyl 2-sulfate (20 and 11.9 mol %), β-d-galactopyranosyl 2-sulfate 4,6-O-(1′-carboxyethylidene) (8.9 and 6.0 mol %) and β-d-galactopyranosyl 2,6-sulfate (5.4 and 18.6 mol %), together with 3,6-anhydro-α-l-galactopyranosyl (11.4 and 7.3 mol %) and 3,6-anhydro-α-l-galactopyranosyl 2-sulfate (4.9 and 15.4 mol %) and minor quantities of 12-15 other structural units.Preparative alkaline treatment carried out on fraction (S2S-3) produced a quantitative formation of 3,6-anhydro α-l-galactopyranosyl units from precursor units (α-l-galactose 6-sulfate and α-l-galactose 2,6-sulfate). Kinetic studies on this 3,6-anhydro cyclization show a rate constant of 5.2 × 104 s−1 indicating diads of the type G→L6S/2,6S. Data from chemical, spectroscopic and kinetic studies suggest that, in S2S-3, the agaran block in the d,l-hybrid galactan is composed of the following diads: G(6R)→L6S/2,6S and G2S(P)(2,6S)→LA(2S)(2R)(2M) and the carrageenan block of G2S(P)→D(2S)(2,3S)(3S)(3,6S) in a molar ratio of agaran to carrageenan structures of ∼2:1.  相似文献   

9.
We used small-angle X-ray solution scattering (SAXS) technique to investigate the nucleotide-mediated conformational changes of the head domains [subfragment 1 (S1)] of myosin V and VI processive motors that govern their directional preference for motility on actin. Recombinant myosin V-S1 with two IQ motifs (MV-S1IQ2) and myosin VI-S1 (MVI-S1) were engineered from Sf9 cells using a baculovirus expression system. The radii of gyration (Rg) of nucleotide-free MV-S1IQ2 and MVI-S1 were 48.6 and 48.8 Å, respectively. In the presence of ATP, the Rg value of MV-S1IQ2 decreased to 46.7 Å, while that of MVI-S1 increased to 51.7 Å, and the maximum chord length of the molecule decreased by ca 9% for MV-S1IQ2 and increased by ca 6% for MVI-S1. These opposite directional changes were consistent with those occurring in S1s with ADP and Vi or AlF4− 2 bound (i.e., in states mimicking the ADP/Pi-bound state of ATP hydrolysis). Binding of AMPPNP induced Rg changes of both constructs similar to those in the presence of ATP, suggesting that the timing of the structural changes for their motion on actin is upon binding of ATP (the pre-hydrolysis state) during the ATPase cycle. Binding of ADP to MV-S1IQ2 and MVI-S1 caused their Rg values to drop below those in the nucleotide-free state. Thus, upon the release of Pi, the reverse conformational change could occur, coupling to drive the directional motion on actin. The amount of Rg change upon the release of Pi was ca 6.4 times greater in MVI-S1 than in MV-S1IQ2, relating to the production of the large stroke of the MVI motor during its translocation on actin. Atomic structural models for these S1s based upon the ab initio shape reconstruction from X-ray scattering data were constructed, showing that MVI-S1 has the light-chain-binding domain positioned in the opposite direction to MV-S1IQ2 in both the pre- and post-powerstroke transition. The angular change between the light chain-binding domains of MV-S1IQ2 in the pre- to post-powerstroke transition was ∼ 50°, comparable to that of MII-S1. On the other hand, that of MVI-S1 was ∼ 100° or ∼ 130° much less than the currently postulated changes to allow the maximal stroke size of myosin VI-S1 but still significantly larger than those of other myosins reported so far. The results suggest that some additional alterations or elements are required for MVI-S1 to take maximal working strokes along the actin filament.  相似文献   

10.
The four classes of heterotrimeric G-protein α subunits act as molecular routers inside cells, gating signals based on a bound guanosine nucleotide (guanosine 5′-triphosphate versus guanosine 5′-diphosphate). Ligands that specifically target individual subunits provide new tools for monitoring and modulating these networks, but are challenging to design due to the high sequence homology and structural plasticity of the Gα-binding surface. Here we have created an mRNA display library of peptides based on the short Gα-modulating peptide R6A-1 and selected variants that target a convergent protein-binding surface of Gαs·guanosine 5′-diphosphate. After selection/evolution, the most Gαs-specific peptide, Gαs(s)-binding peptide (GSP), was used to design a second-generation library, resulting in several new affinity- and selectivity-matured peptides denoted as mGSPs. The two-step evolutionary walk from R6A-1 to mGSP-1 resulted in an 8000-fold inversion in binding specificity, altered seven out of nine residues in the starting peptide core, and incorporated both positive and negative design steps. The resulting mGSP-1 peptide shows remarkable selectivity and affinity, exhibiting little or no binding to nine homologous Gα subunits or human H-Ras, and even discriminates the Gαs splice variant Gαs(l). Selected peptides make specific contacts with the effector-binding region of Gα, which may explain an interesting bifunctional activity observed in GSP. Overall, our work demonstrates a design of simple, linear, highly specific peptides that target a protein-binding surface of Gαs and argues that mRNA display-based selection/evolution is a powerful route for targeting protein families with high class specificity and state specificity.  相似文献   

11.
Escherichia coli dihydrofolate reductase (DHFR) catalyzes the reduction of dihydrofolate to tetrahydrofolate. During the catalytic cycle, DHFR undergoes conformational transitions between the closed (CS) and occluded (OS) states that, respectively, describe whether the active site is closed or occluded by the Met20 loop. The CS→OS and the reverse transition may be viewed as allosteric transitions. Using a sequence-based approach, we identify a network of residues that represents the allostery wiring diagram. Many of the residues in the allostery wiring diagram, which are dispersed throughout the adenosine-binding domain as well as the loop domain, are not conserved. Several of the residues in the network have been previously shown by NMR experiments, mutational studies, and molecular dynamics simulations to be linked to equilibration conformational fluctuations of DHFR. To further probe the nature of events that occur during conformational fluctuations, we use a self-organized polymer model to monitor the kinetics of the CS→OS and the reverse transitions. During the CS→OS transition, coordinated changes in a number of residues in the loop domain enable the Met20 loop to slide along the α-helix in the adenosine-binding domain. Sliding is triggered by pulling of the Met20 loop by the βG-βH loop and the pushing action of the βG-βH loop. The residues that facilitate the Met20 loop motion are part of the network of residues that transmit allosteric signals during the CS→OS transition. Replacement of M16 and G121, whose Cα atoms are about 4.3 Å in the CS, by a disulfide cross-link impedes that CS→OS transition. The order of events in the OS→CS transition is not the reverse of the forward transition. The contact Glu18-Ser49 in the OS persists until the sliding of the Met20 loop is nearly complete. The ensemble of structures in the transition state in both the allosteric transitions is heterogeneous. The most probable transition-state structure resembles the OS (CS) in the CS→OS (OS→CS) transition, which is in accord with the Hammond postulate. Structures resembling the OS (CS) are present as minor (∼ 1-3%) components in equilibrated CS (OS) structures.  相似文献   

12.
Transketolase from Saccharomyces cerevisiae exhibits a rarely reported activity with a methylated analogue of the native cofactor, 4′-methylamino-thiamin diphosphate. We demonstrated the kinetic stability of the dihydroxyethyl carbanion/enamine intermediate to be dependent on the functionality of the 4′-aminopyrimidine moiety of thiamin diphosphate [R. Golbik, L.E. Meshalkina, T. Sandalova, K. Tittmann, E. Fiedler, H. Neef, S. König, R. Kluger, G.A. Kochetov, G. Schneider, G. Hübner, Effect of coenzyme modification on the structural and catalytic properties of wild-type transketolase and of the variant E418A from Saccharomyces cerevisae, FEBS J. (2005) 272 1326-1342]. This paper extends these investigations of the function of the coenzyme’s aminopyrimidine in transketolase catalysis exemplified for the 4′-monomethylamino-thiamin diphosphate analogue. Here, we report near UV circular dichroism data and NMR-based analysis of reaction intermediates that give evidence for a strong destabilisation of the carbanion/enamine of DHE-4’-monomethylamino-thiamin diphosphate on the enzyme. A new negative band in near UV circular dichroism arising during turnover is attributed to the conjugate acid of the carbanion/enamine intermediate, an assignment additionally corroborated by 1H NMR-based intermediate analysis. As opposed to the kinetically stabilized carbanion/enamine intermediate in transketolase when reconstituted with the native cofactor, DHE-4′-monomethylamino-thiamin diphosphate is rapidly released from the active centers during turnover and accumulates in the medium on a preparative scale.  相似文献   

13.
Here, we present a study of the conformational changes of the quaternary structure of Escherichia coli aspartate transcarbamoylase, as monitored by time-resolved small-angle X-ray scattering, upon combining with substrates, substrate analogs, and nucleotide effectors at temperatures between 5 and 22 °C, obviating the need for ethylene glycol. Time-resolved small-angle X-ray scattering time courses tracking the T → R structural change after mixing with substrates or substrate analogs appeared to be a single phase under some conditions and biphasic under other conditions, which we ascribe to multiple ligation states producing a time course composed of multiple rates. Increasing the concentration of substrates up to a certain point increased the T → R transition rate, with no further increase in rate beyond that point. Most strikingly, after addition of N-phosphonacetyl-l-aspartate to the enzyme, the transition rate was more than 1 order of magnitude slower than with the natural substrates. These results on the homotropic mechanism are consistent with a concerted transition between structural and functional states of either low affinity, low activity or high affinity, high activity for aspartate. Addition of ATP along with the substrates increased the rate of the transition from the T to the R state and also decreased the duration of the R-state steady-state phase. Addition of CTP or the combination of CTP/UTP to the substrates significantly decreased the rate of the T → R transition and caused a shift in the enzyme population towards the T state even at saturating substrate concentrations. These results on the heterotropic mechanism suggest a destabilization of the T state by ATP and a destabilization of the R state by CTP and CTP/UTP, consistent with the T and R state crystallographic structures of aspartate transcarbamoylase in the presence of the heterotropic effectors.  相似文献   

14.
The ratio of the internucleotide dipolar coupling and the corresponding one-bond imino 15N-1H dipolar coupling provides a measure for the N···H/H-N distance ratio. Measurements were carried out for a dodecamer, d(CGCGAATTCGCG)2, in which a C-G and an A-T basepair were uniformly enriched in 15N. When assuming H-bonds to be perfectly linear, dipolar data indicate time-averaged hydrogen bond lengths of 1.80±0.03 Å for A-T and 1.86±0.02 Å for C-G. When using H-bond orientations from high resolution X-ray data, H-bond lengths are about 0.1 Å shorter.  相似文献   

15.
Determining the network of residues that transmit allosteric signals is crucial to understanding the function of biological nanomachines. During the course of a reaction cycle, biological machines in general, and Escherichia coli chaperonin GroEL in particular, undergo large-scale conformational changes in response to ligand binding. Normal mode analyses, based on structure-based coarse-grained models where each residue is represented by an α carbon atom, have been widely used to describe the motions encoded in the structures of proteins. Here, we propose a new Cα-side chain elastic network model of proteins that includes information about the physical identity of each residue and accurately accounts for the side-chain topology and packing within the structure. Using the Cα-side chain elastic network model and the structural perturbation method, which probes the response of a local perturbation at a given site at all other sites in the structure, we determine the network of key residues (allostery wiring diagram) responsible for the T → R and R″ → T transitions in GroEL. A number of residues, both within a subunit and at the interface of two adjacent subunits, are found to be at the origin of the positive cooperativity in the ATP-driven T → R transition. Of particular note are residues G244, R58, D83, E209, and K327. Of these, R38, D83, and K327 are highly conserved. G244 is located in the apical domain at the interface between two subunits; E209 and K327 are located in the apical domain, toward the center of a subunit; R58 and D83 are equatorial domain residues. The allostery wiring diagram shows that the network of residues are interspersed throughout the structure. Residues D83, V174, E191, and D359 play a critical role in the R″ → T transition, which implies that mutations of these residues would compromise the ATPase activity. D83 and E191 are also highly conserved; D359 is moderately conserved. The negative cooperativity between the rings in the R″ → T transition is orchestrated through several interface residues within a single ring, including N10, E434, D435, and E451. Signal from the trans ring that is transmitted across the interface between the equatorial domains is responsible for the R″ → T transition. The cochaperonin GroES plays a passive role in the R″ → T transition. Remarkably, the binding affinity of GroES for GroEL is allosterically linked to GroEL residues 350-365 that span helices K and L. The movements of helices K and L alter the polarity of the cavity throughout the GroEL functional cycle and undergo large-scale motions that are anticorrelated with the other apical domain residues. The allostery wiring diagrams for the T → R and R″ → T transitions of GroEL provide a microscopic foundation for the cooperativity (anticooperativity) within (between) the ring (rings). Using statistical coupling analysis, we extract evolutionarily linked clusters of residues in GroEL and GroES. We find that several substrate protein binding residues as well as sites related to ATPase activity belong to a single functional network in GroEL. For GroES, the mobile loop residues and GroES/GroES interface residues are linked.  相似文献   

16.
By means of time-resolved electron paramagnetic resonance (EPR) spectroscopy, the photoexcited structural changes of site-directed spin-labeled bacteriorhodopsin are studied. A complete set of cysteine mutants of the C-D loop, positions 100-107, and of the E-F loop, including the first alpha-helical turns of helices E and F, positions 154-171, was modified with a methanethiosulfonate spin label. The EPR spectral changes occurring during the photocycle are consistent with a small movement of helix C and an outward tilt of helix F. These helix movements are accompanied by a rearrangement of the E-F loop and of the C-terminal turn of helix E. The kinetic analysis of the transient EPR data and the absorbance changes in the visible spectrum reveals that the conformational change occurs during the lifetime of the M intermediate. Prominent rearrangements of nitroxide side chains in the vicinity of D96 may indicate the preparation of the reprotonation of the Schiff base. All structural changes reverse with the recovery of the bacteriorhodopsin initial state.  相似文献   

17.
At room temperature, the chlorophyll (Chl) a fluorescence induction (FI) kinetics of plants, algae and cyanobacteria go through two maxima, P at ∼ 0.2-1 and M at ∼ 100-500 s, with a minimum S at ∼ 2-10 s in between. Thus, the whole FI kinetic pattern comprises a fast OPS transient (with O denoting origin) and a slower SMT transient (with T denoting terminal state). Here, we examined the phenomenology and the etiology of the SMT transient of the phycobilisome (PBS)-containing cyanobacterium Synechococcus sp PCC 7942 by modifying PBS → Photosystem (PS) II excitation transfer indirectly, either by blocking or by maximizing the PBS → PS I excitation transfer. Blocking the PBS → PS I excitation transfer route with N-ethyl-maleimide [NEM; A. N. Glazer, Y. Gindt, C. F. Chan, and K.Sauer, Photosynth. Research 40 (1994) 167-173] increases both the PBS excitation share of PS II and Chl a fluorescence. Maximizing it, on the other hand, by suspending cyanobactrial cells in hyper-osmotic media [G. C. Papageorgiou, A. Alygizaki-Zorba, Biochim. Biophys. Acta 1335 (1997) 1-4] diminishes both the PBS excitation share of PS II and Chl a fluorescence. Here, we show for the first time that, in either case, the slow SMT transient of FI disappears and is replaced by continuous P → T fluorescence decay, reminiscent of the typical P → T fluorescence decay of higher plants and algae. A similar P → T decay was also displayed by DCMU-treated Synechococcus cells at 2 °C. To interpret this phenomenology, we assume that after dark adaptation cyanobacteria exist in a low fluorescence state (state 2) and transit to a high fluorescence state (state 1) when, upon light acclimation, PS I is forced to run faster than PS II. In these organisms, a state 2 → 1 fluorescence increase plus electron transport-dependent dequenching processes dominate the SM rise and maximal fluorescence output is at M which lies above the P maximum of the fast FI transient. In contrast, dark-adapted plants and algae exist in state 1 and upon illumination they display an extended P → T decay that sometimes is interrupted by a shallow SMT transient, with M below P. This decay is dominated by a state 1 → 2 fluorescence lowering, as well as by electron transport-dependent quenching processes. When the regulation of the PBS → PS I electronic excitation transfer is eliminated (as for example in hyper-osmotic suspensions, after NEM treatment and at low temperature), the FI pattern of Synechococcus becomes plant-like.  相似文献   

18.
Adenylyl cyclases (ACs) belonging to three nonhomologous classes (II, III, and IV) have been structurally characterized, enabling a comparison of the mechanisms of cyclic adenosine 3′,5′-monophosphate biosynthesis. We report the crystal structures of three active-site complexes for Yersinia pestis class IV AC (AC-IV)—two with substrate analogs and one with product. Mn2+ binds to all three phosphates, and to Glu12 and Glu136. Electropositive residues Lys14, Arg63, Lys76, Lys111, and Arg113 also form hydrogen bonds to phosphates. The conformation of the analogs is suitable for in-line nucleophilic attack by the ribose O3′ on α-phosphate (distance ∼ 4 Å). In the product complex, a second Mn ion is observed to be coordinated to both ribose 2′ oxygen and ribose 3′ oxygen. Observation of both metal sites, together with kinetic measurements, provides strong support for a two-cation mechanism. Eleven active-site mutants were also made and kinetically characterized. These findings and comparisons with class II and class III enzymes enable a detailed transphyletic analysis of the AC mechanism. Consistent with its lack of coordination to purine, Y. pestis AC-IV cyclizes both ATP and GTP. As in other classes of AC, the ribose is loosely bound, and as in class III, no base appears to ionize the O3′ nucleophile. Different syn/anti conformations suggest that the mechanism involves a conformational transition, and further evidence suggests a role for ribosyl pseudorotation. With resolutions of 1.6-1.7 Å, these are the most detailed active-site ligand complexes for any class of this ubiquitous signaling enzyme.  相似文献   

19.
In the P-domain, the 369-DKTGTLT and the 709-GDGVNDSPALKK segment are highly conserved during evolution of P-type E1–E2-ATPase pumps irrespective of their cation specificities. The focus of this article is on evaluation of the role of the amino acid residues in the P domain of the subunit of Na,K-ATPase for the E1P[3Na] E2P[2Na] conversion, the K+-activated dephosphorylation, and the transmission of these changes to and from the cation binding sites. Mutations of residues in the TGDGVND loop show that Asp710 is essential, and Asn713 is important, for Mg2+ binding and formation of the high-energy MgE1P[3Na] intermediate. In contrast Asp710 and Asp713 do not contribute to Mg2+ binding in the E2P–ouabain complex. Transition to E2P thus involves a shift of Mg2+ coordination away from Asp710 and Asn713 and the two residues become more important for K+-activated hydrolysis of the acyl phosphate bond at Asp369. Transmission of structural changes between the P-domain and cation sites in the membrane domain is evaluated in light of the protein structure, and the information from proteolytic or metal-catalyzed cleavage and mutagenesis studies.  相似文献   

20.
The flavivirus 2′-O-nucleoside N-terminal RNA methyltransferase (MTase) enzyme is responsible for methylating the viral RNA cap structure. To increase our understanding of the mechanism of viral RNA cap binding we performed a detailed structural and biochemical characterization of the guanosine cap-binding pocket of the dengue (DEN) and yellow fever (YF) virus MTase enzymes. We solved an improved 2.1 Å resolution crystal structure of DEN2 Mtase, new 1.5 Å resolution crystal structures of the YF virus MTase domain in apo form, and a new 1.45 Å structure in complex with guanosine triphosphate and RNA cap analog. Our structures clarify the previously reported DEN MTase structure, suggest novel protein-cap interactions, and provide a detailed view of guanine specificity. Furthermore, the structures of the DEN and YF proteins are essentially identical, indicating a large degree of structural conservation amongst the flavivirus MTases. Guanosine triphosphate analog competition assays and mutagenesis analysis, performed to analyze the biochemical characteristics of cap binding, determined that the major interaction points are (i) guanine ring via π−π stacking with Phe24, N1 hydrogen interaction with the Leu19 backbone carbonyl via a water bridge, and C2 amine interaction with Leu16 and Leu19 backbone carbonyls; (ii) ribose 2′ hydroxyl interaction with Lys13 and Asn17; and (iii) α-phosphate interactions with Lys28 and Ser215. Based on our mutational and analog studies, the guanine ring and α-phosphate interactions provide most of the energy for cap binding, while the combination of the water bridge between the guanine N1 and Leu19 carbonyl and the hydrogen bonds between the C2 amine and Leu16/Leu19 carbonyl groups provide for specific guanine recognition. A detailed model of how the flavivirus MTase protein binds RNA cap structures is presented.  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