首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
The exciton transfer between light-harvesting complex 1(LH1) and photosynthetic reaction center dimer is investigated theoretically. We assume a ring shape structure of the LH1 complex with dimer in the ring centre. The kinetic equations which describe the energy transfer between the antenna complex and reaction center dimer were derived. It was shown that the dimer does not act as a photon trap. There is a weak localization of the exciton on the dimer and there is relatively rapid back exciton transfer from dimer to antenna complex which depends on the number of the pigment molecules in the antenna ring. The relation between the rates of the exciton transfer from the antenna complex to dimer and back transfer from dimer to antenna complex has been derived.  相似文献   

2.
Perfusion of isolated small intestine with the procyanidin dimers B2 and B5 extracted from cocoa indicated that both forms of dimer are transferred to the serosal side of enterocytes but only to a very small extent (<1% of the total transferred flavanol-like compounds). However, perfusion of dimer mainly resulted in large amounts of unmetabolised/unconjugated epicatechin monomer being detected on the serosal side (95.8%). The cleavage of dimer during transfer seemed to be energy-dependent, requiring an intact cell system, as incubation with jejunal homogenates failed to yield epicatechin. Low levels methylated dimer were also detected (3.2%), but no conjugates and metabolites of epicatechin indicating that metabolism of monomer and dimer is limited during dimer cleavage/translocation. The methylation of dimer may be by catechol-O-methyltransferase, however, at high concentrations of dimer COMT activity is reduced leading to an inhibition of both monomer and dimer O-methylation.  相似文献   

3.
The binding of EGF induces dimerization of its receptor, leading to the stimulation of its intracellular tyrosine kinase activity. Kinase activation occurs within the context of an asymmetric dimer in which one kinase domain serves as the activator for the other kinase domain but is not itself activated. How ligand binding is related to the formation and dynamics of this asymmetric dimer is not known. The binding of EGF to its receptor is negatively cooperative--that is, EGF binds with lower affinity to the second site on the dimer than to the first site on the dimer. In this study, we analyzed the binding of (125)I-EGF to a series of EGF receptor mutants in the intracellular juxtamembrane domain and demonstrate that the most membrane-proximal portion of this region plays a significant role in the genesis of negative cooperativity in the EGF receptor. The data are consistent with a model in which the binding of EGF to the first site on the dimer induces the formation of one asymmetric kinase dimer. The binding of EGF to the second site is required to disrupt the initial asymmetric dimer and allow the formation of the reciprocal asymmetric dimer. Thus, some of the energy of binding to the second site is used to reorient the first asymmetric dimer, leading to a lower binding affinity and the observed negative cooperativity.  相似文献   

4.
We carried out molecular dynamics simulations and free energy calculations for a series of ternary and diplex models for the HipA protein, HipB dimer, and DNA molecule to address the mechanism of HipA sequestration and the binding order of events from apo HipB/HipA to 2HipA + HipB dimer + DNA complex. The results revealed that the combination of DNA with the HipB dimer is energetically favorable for the combination of HipB dimer with HipA protein. The binding of DNA to HipB dimer induces a long‐range allosteric communication from the HipB2‐DNA interface to the HipA–HipB2 interface, which involves the closeness of α1 helices of HipB dimer to HipA protein and formations of extra hydrogen bonds in the HipA–HipB2 interface through the extension of α2/3 helices in the HipB dimer. These simulated results suggested that the DNA molecule, as a regulative media, modulates the HipB dimer conformation, consequently increasing the interactions of HipB dimer with the HipA proteins, which explains the mechanism of HipA sequestration reported by the previous experiment. Simultaneously, these simulations also explored that the thermodynamic binding order in a simulated physiological environment, that is, the HipB dimer first bind to DNA to form HipB dimer + DNA complex, then capturing strongly the HipA proteins to form a ternary complex, 2HipA + HipB dimer + DNA, for sequestrating HipA in the nucleoid. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

5.
DnaK, a major Hsp70 molecular chaperones in Escherichia coli, is a widely used model for studying Hsp70s. We recently solved a crystal structure of DnaK in complex with ATP and showed that DnaK was packed as a dimer in the crystal structure. Our previous biochemical studies supported the formation of a specific DnaK dimer as observed in the crystal structure in solution in the presence of ATP and suggested an important role of this dimer in efficient interaction with Hsp40 co-chaperones. In this study, we dissected the biochemical properties of this DnaK dimer. To restrict DnaK in this dimer form, we mutated two residues on the dimer interface to cysteine, A303C, and H541C. Upon oxidation, this DnaK-A303C-H541C protein formed a specific dimer linked by disulfide bonds formed between A303C and H541C only in the presence of ATP, consistent with the crystal structure. Intriguingly, this disulfide-bond-linked dimer of DnaK-A303C-H541C has reduced ATPase activity and decreased affinity for peptide substrate. More interestingly, unlike wild-type DnaK, the peptide substrate-binding kinetics of this dimer is drastically accelerated even in the absence of ATP, suggesting this dimer is restricted in an ATP-bound conformation regardless of nucleotide bound, which was further supported by our analysis using tryptophan fluorescence and ATP-induced peptide release. Thus, formation of the dimer restricted DnaK in an ATP-bound state and blocked the progression through the chaperone cycle. Productive progression through the chaperone cycle requires the dissociation of this transient dimer. Surprisingly, a significantly compromised interaction with Hsp40 co-chaperone was observed for this disulfide-bond-linked dimer. Thus, dissociation of this DnaK dimer is equally crucial for efficient Hsp40 interaction. An initial interaction between Hsp70 and Hsp40 requires the formation of DnaK dimer; but a stable Hsp70-Hsp40 interaction may follow the dissociation of the dimer.  相似文献   

6.
The extracellular domains of cadherins are known to play a major role in cell adhesion, although the structures involved in this process remain unclear. We have used molecular dynamics to characterize the conformational and thermodynamic properties of two of the dimer interfaces identified in E-cadherin crystals and involving the two outermost exodomains (EC1 and EC2): a dimer involving exchange of the N-terminal strand (referred to as the "swapped" dimer) and a "staggered" dimer involving an EC1-EC2 interface. The results show that the staggered dimer involves a much smaller interface area and is notably less stable than the swapped dimer. It is also found that, despite its stability, the swapped dimer undergoes a conformational transition leading to a structure closer to that experimentally observed for the homologous C-cadherin. Finally, comparing the simulated dimer structures with the sequences of E-, C-, and N-cadherins shows that the swapped dimer interface involves surprisingly few residues that vary from family to family and notably no changes between the E- and C-cadherin exodomains.  相似文献   

7.
8.
Pallesen J 《Biochemistry》2011,50(28):6170-6177
The formation of a genomic RNA dimer is critical for the HIV-1 replication cycle, and dimerization is known to initiate within the 5'UTR (5' untranslated region) of the viral RNA. However, the 5'UTR constitutes the 335 terminal nucleotides, and because of this considerable size, it has been difficult to study the global structure using conventional structural methods. Here, the atomic force microscope has been used to directly visualize the dimer formed from RNAs including HIV-1 nucleotides 1-744. Gold nanocolloids were deposited on the primer binding site regions in the dimer as an internal control. The dimer showed distinct ring morphology with up to two gold nanocolloids deposited within the ring and one or two strands extending from the ring. This morphology implies a dimer including a DIS-DIS (dimerization initiation site)-containing 3' dimer linkage site (DLS) and a TAR-TAR (trans-activation region)-containing 5'DLS. Furthermore, the dimer was formed under the influence of Mg(2+) and was imaged with an atomic force microscope under buffer conditions. The overall ring morphology containing a 5'DLS and a 3'DLS with one or two strands extending from it was conserved in these atomic force microscopy images. This indicates that the observed dimer morphology is physiologically significant. Moreover, evidence of multiple dimer interstrand contacts downstream of the major splice donor were observed, which indicates a component in the selection of full-length genomic RNA in dimer formation during virion packaging.  相似文献   

9.
UV RESISTANCE LOCUS8 (UVR8) is a photoreceptor for ultraviolet‐B (UV‐B) light that initiates photomorphogenic responses in plants. UV‐B photoreception causes rapid dissociation of dimeric UVR8 into monomers that interact with CONSTITUTIVELY PHOTOMORPHOGENIC1 (COP1) to initiate signal transduction. Experiments with purified UVR8 show that the dimer is maintained by salt‐bridge interactions between specific charged amino acids across the dimer interface. However, little is known about the importance of these charged amino acids in determining dimer/monomer status and UVR8 function in plants. Here we evaluate the use of different methods to examine dimer/monomer status of UVR8 and show that mutations of several salt‐bridge amino acids affect dimer/monomer status, interaction with COP1 and photoreceptor function of UVR8 in vivo. In particular, the salt‐bridges formed between arginine 286 and aspartates 96 and 107 are key to dimer formation. Mutation of arginine 286 to alanine impairs dimer formation, interaction with COP1 and function in vivo, whereas mutation to lysine gives a weakened dimer that is functional in vivo, indicating the importance of the positive charge of the arginine/lysine residue for dimer formation. Notably, a UVR8 mutant in which aspartates 96 and 107 are conservatively mutated to asparagine is strongly impaired in dimer formation but mediates UV‐B responses in vivo with a similar dose–response relationship to wild‐type. The UV‐B responsiveness of this mutant does not correlate with dimer formation and monomerisation, indicating that monomeric UVR8 has the potential for UV‐B photoreception, initiating signal transduction and responses in plants.  相似文献   

10.
Stokeley D  Bemporad D  Gavaghan D  Sansom MS 《Biochemistry》2007,46(47):13573-13580
Saposin B is a water soluble alpha-helical protein which can bind to membranes and extract selected lipids, especially cerebroside sulfates. The X-ray structure of saposin B is homodimeric. There are two conformations of the dimer in the crystal-one with a closed central cavity (the AB dimer) and one (the CD dimer) with a more open cavity. We have conducted a series of short (5 ns) molecular dynamics simulations of saposin B, starting from both the AB and CD conformations and with/without bound lipid and/or water molecules within the central hydrophobic cavity. The more open (CD) dimer showed greater conformational drift than the AB dimer. The conformational drift was also somewhat higher in the absence of bound lipid. Two more extended (30 ns) simulations of AB and CD dimers were performed and analyzed in terms of changes in intersubunit packing within the dimers. The AB dimer remained largely unchanged in conformation over the duration of the extended simulation. In contrast, the CD dimer underwent a substantial conformational change corresponding to a 'scissor' motion of the two monomers so as to compress the central cavity to a more closed conformation than that seen in the AB dimer structure. A H-bond between the Q53 and Y54 side chains of the alpha3 helices of the two opposing monomers seems to hold the dimer in this 'scissor-closed' conformation. We suggest that a cycle of conformational changes, expanding and compressing the central cavity of the saposin B dimer, may play a key role in facilitating lipid extraction from bilayers.  相似文献   

11.
Stefin A folds as a monomer under strongly native conditions. We have observed that under partially denaturing conditions in the temperature range from 74 to 93 degrees C it folds into a dimer, while it is monomeric above the melting temperature of 95 degrees C. Below 74 degrees C the dimer is trapped and it does not dissociate. The dimer is a folded and structured protein as judged by CD and NMR, nevertheless it is no more functional as an inhibitor of cysteine proteases. The monomer-dimer transition proceeds at a slow rate and the activation energy of dimerization at 99 kcal/mol is comparable to the unfolding enthalpy. A large and negative dimerization enthalpy of -111(+/- 8) kcal/mol was calculated from the temperature dependence of the dissociation constant. An irreversible pretransition at 10-15 deg. below the global unfolding temperature has been observed previously by DSC and can now be assigned to the monomer-dimer transition. Backbone resonances of all the dimer residues were assigned using 15N isotopically enriched protein. The dimer is symmetric and the chemical shift differences between the monomer and dimer are localized around the tripartite hydrophobic wedge, which otherwise interacts with cysteine proteases. Hydrogen exchange protection factors of the residues affected by dimer formation are higher in the dimer than in the monomer. The monomer to dimer transition is accompanied by a rapid exchange of all of the amide protons which are protected in the dimer, indicating that the transition state is unfolded to a large extent. Our results demonstrate that the native monomeric state of stefin A is actually metastable but is favored by the kinetics of folding. The substantial energy barrier which separates the monomer from the more stable dimer traps each state under native conditions.  相似文献   

12.
Acharya A  Ruvinov SB  Gal J  Moll JR  Vinson C 《Biochemistry》2002,41(48):14122-14131
We use a heterodimerizing leucine zipper system to examine the contribution of the interhelical a-a' interaction to dimer stability for six amino acids (A, V, L, I, K, and N). Circular dichroism (CD) spectroscopy monitored the thermal denaturation of 36 heterodimers that generate six homotypic and 30 heterotypic a-a' interactions. Isoleucine (I-I) is the most stable homotypic a-a' interaction, being 9.2 kcal/mol per dimer more stable than the A-A interaction and 4.0 kcal/mol per dimer more stable than either the L-L or V-V interaction, and 7.0 kcal/mol per dimer more stable than the N-N interaction. Only lysine was less stable than alanine. An alanine-based double-mutant thermodynamic cycle calculated coupling energies between the a and a' positions in the heterodimer. The aliphatic amino acids L, V, and I prefer to form homotypic interactions with coupling energies of -0.6 to -0.9 kcal/mol per dimer, but the heterotypic aliphatic interactions have positive coupling energies of <1.0 kcal/mol per dimer. The asparagine homotypic interaction has a coupling energy of -0.5 kcal/mol per dimer, while heterotypic interactions with the aliphatic amino acids produce coupling energies ranging from 2.6 to 4.9 kcal/mol per dimer. The homotypic K-K interaction is 2.9 kcal/mol per dimer less stable than the A-A interaction, but the coupling energy is only 0.3 kcal/mol per dimer. Heterotypic interactions with lysine and either asparagine or aliphatic amino acids produce similar coupling energies ranging from -0.2 to -0.7 kcal/mol per dimer. Thus, of the amino acids that were examined, asparagine contributes the most to dimerization specificity because of the large positive coupling energies in heterotypic interactions with the aliphatic amino acids which results in the N-N homotypic interaction.  相似文献   

13.
Retroviral integrases are reported to form alternate dimer assemblies like the core–core dimer and reaching dimer. The core–core dimer is stabilized predominantly by an extensive interface between two catalytic core domains. The reaching dimer is stabilized by N-terminal domains that reach to form intermolecular interfaces with the other subunit’s core and C-terminal domains (CTD), as well as CTD–CTD interactions. In this study, molecular dynamics (MD), Brownian dynamics (BD) simulations, and free energy analyses, were performed to elucidate determinants for the stability of the reaching dimer forms of full-length Avian Sarcoma Virus (ASV) and Human Immunodeficiency Virus (HIV) IN, and to examine the role of the C-tails (the last ~16–18 residues at the C-termini) in their structural dynamics. The dynamics of an HIV reaching dimer derived from small angle X-ray scattering and protein crosslinking data, was compared with the dynamics of a core–core dimer model derived from combining the crystal structures of two-domain fragments. The results showed that the core domains in the ASV reaching dimer express free dynamics, whereas those in the HIV reaching dimer are highly stable. BD simulations suggest a higher rate of association for the HIV core–core dimer than the reaching dimer. The predicted stability of these dimers was therefore ranked in the following order: ASV reaching dimer < HIV reaching dimer < composite core–core dimer. Analyses of MD trajectories have suggested residues that are critical for intermolecular contacts in each reaching dimer. Tests of these predictions and insights gained from these analyses could reveal a potential pathway for the association and dissociation of full-length IN multimers.  相似文献   

14.
The location of the 1:2 borate-diol ester cross-link in the dimer of the plant cell wall polysaccharide rhamnogalacturonan II (RG-II) has been determined. The ester cross-links the apiofuranosyl residue of the 2-O-methyl-D-xylose-containing side chains in each of the subunits of the dimer. The apiofuranosyl residue in each of the two aceric acid-containing side chains is not esterified. The site of borate esterification is identical in naturally occurring and in in vitro synthesized dimer. Pb2+, La3+, and Ca2+ increase dimer formation in vitro in a concentration- and pH-dependent manner. Pb2+ is the most effective cation. The dimer accounts for 55% of the RG-II when the monomer (0.5 mM) is treated for 5 min at pH 3.5 with boric acid (1 mM) and Pb2+ (0.5 mM); at pH 5 the rate of conversion is somewhat slower. Hg2+ does not increase the rate of dimer formation. A cation's charge density and its ability to form a coordination complex with RG-II, in addition to steric factors, may regulate the rate and stability of dimer formation in vitro. Our data provide evidence that the structure of RG-II itself determines which apiofuranosyl residues are esterified with borate and that in the presence of boric acid and certain cations, two RG-II monomers self-assemble to form a dimer.  相似文献   

15.
Xu Y  Moseley JB  Sagot I  Poy F  Pellman D  Goode BL  Eck MJ 《Cell》2004,116(5):711-723
Formin proteins participate in a wide range of cytoskeletal processes in all eukaryotes. The defining feature of formins is a highly conserved approximately 400 residue region, the Formin Homology-2 (FH2) domain, which has recently been found to nucleate actin filaments. Here we report crystal structures of the S. cerevesiae Bni1p FH2 domain. The mostly alpha-helical FH2 domain forms a unique "tethered dimer" in which two elongated actin binding heads are tied together at either end by an unusual lasso and linker structure. Biochemical and crystallographic observations indicate that the dimer is stable but flexible, with flexibility between the two halves of the dimer conferred by the linker segments. Although each half of the dimer is competent to interact with filament ends, the intact dimer is required for actin nucleation and processive capping. The tethered dimer architecture may allow formins to stair-step on the barbed end of an elongating nascent filament.  相似文献   

16.
Interleukin-8 (IL-8), a member of the chemokine superfamily, exists as both monomers and dimers, and mediates its function by binding to neutrophil CXCR1 and CXCR2 receptors that belong to the G protein-coupled receptor class. It is now well established that the monomer functions as a high-affinity ligand, but the binding affinity of the dimer remains controversial. The approximately 1000-fold difference between monomer-dimer equilibrium constant (microM) and receptor binding constant (nM) of IL-8 does not allow receptor-binding affinity measurements of the native IL-8 dimer. In this study, we overcame this roadblock by creating a "trapped" nondissociating dimer that contains a disulfide bond across the dimer interface at the 2-fold symmetry point. The NMR studies show that the structure of this trapped dimer is indistinguishable from the native dimer. The trapped dimer, compared to a trapped monomer, bound CXCR1 with approximately 70-fold and CXCR2 with approximately 20-fold lower affinities. Receptor binding involves two interactions, between the IL-8 N-loop and receptor N-domain residues, and between IL-8 N-terminal and receptor extracellular loop residues. In contrast to a trapped monomer that bound an isolated CXCR1 N-domain peptide with microM affinity, the trapped dimer failed to show any binding, indicating that dimerization predominantly perturbs the binding of only the N-loop residues. These results demonstrate that only the monomer is a high-affinity ligand for both receptors, and also provide a structural basis for the lower binding affinity of the dimer.  相似文献   

17.
We have constructed a disulfide dimer of S118C azurin, in which two copper centers are coupled through a relatively short covalent pathway, and studied its electron transfer properties. The dimer exhibits intriguing mechanistic properties. Due to the strain in the molecule, caused by the limited accessibility of Cys118, anti-cooperativity occurs in the two step oxidation of the dimer with a difference in redox potential between the two half reactions of 33 mV. Upon oxidation, the dimer favours the semi-reduced over the fully oxidized state, as the Cu(I) site in the semi-reduced dimer is able to stabilize the strained dimer complex. The internal electron transfer is surprisingly slow, which could be partially due to an increase in reorganization energy.  相似文献   

18.
The structure of Drosophila LC8 pH-induced monomer has been determined by NMR spectroscopy using the program AutoStructure. The structure at pH 3 and 30 degrees C is similar to the individual subunits of mammalian LC8 dimer with the exception that a beta strand, which crosses between monomers to form an intersubunit beta-sheet in the dimer, is a flexible loop with turnlike conformations in the monomer. Increased flexibility in the interface region relative to the rest of the protein is confirmed by dynamic measurements based on (15)N relaxation. Comparison of the monomer and dimer structures indicates that LC8 is not a domain swapped dimer.  相似文献   

19.
Refolding a disulfide dimer of cytochrome c   总被引:1,自引:0,他引:1  
A covalent dimer of Saccharomyces cerevisiae iso-1 cytochrome c is stabilized by an interchain disulfide bond involving the cysteine residue penultimate to the C-terminus. The individual chains in the dimer appear to retain the tertiary structural features characteristic for monomeric cytochrome c albeit with some perturbation. The dimer is reversibly denatured by heat, urea, or guanidine hydrochloride in a single cooperative transition whose midpoint is less than that of the monomeric protein. The kinetic profile observed for the refolding of the denatured dimer is characteristic for monomeric cytochromes except for a markedly enhanced slow-phase amplitude.  相似文献   

20.
We explored the stability of the dengue virus envelope (E) protein dimer since it is widely assumed that the E protein dimer is stabilized by drug ligands or antibodies in an acidic environment, neutralizing the virus's ability to fuse with human cells. During this process, a large conformational change of the E protein dimer is required. We performed Molecular Dynamics simulations to mimic the conformational change and stability of the dimer in neutral and acidic conditions with the well-tempered metadynamics method. Furthermore, as a few neutralizing antibodies discovered from dengue patients were reported, we used the same simulation method to examine the influence of a selected antibody on the dimer stability in both neutral and acidic conditions. We also investigated the antibody's influence on a point-mutated E protein that had been reported to interrupt the protein-antibody interaction and result in more than 95% loss of the antibody's binding ability. Our simulation results are highly consistent with the experimental conclusion that binding of the antibody to the E protein dimer neutralizes the virus, especially in a low pH condition, while the mutation of W101A or N153A significantly reduces the antibody's ability in stabilizing the E protein dimer. We demonstrate that well-tempered metadynamics can be used to accurately explore the antibody's interaction on large protein complexes such as the E protein dimer, and the computational approach in this work is promising in future antibody development.  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