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 0 毫秒
1.
Changing the helical propensity of a polypeptide sequence might be expected to affect the conformational properties of the denatured state of a protein. To test this hypothesis, alanines at positions 83 and 87 near the center of helix 3 of cytochrome c′ from Rhodopseudomonas palustris were mutated to serine to decrease the stability of this helix. A set of 13 single histidine variants in the A83S/A87S background were prepared to permit assessment of the conformational properties of the denatured state using histidine-loop formation in 3 M guanidine hydrochloride. The data are compared with previous histidine-heme loop formation data for wild-type cytochrome c′. As expected, destabilization of helix 3 decreases the global stabilities of the histidine variants in the A83S/A87S background relative to the wild-type background. Loop stability versus loop size data yields a scaling exponent of 2.1 ± 0.2, similar to the value of 2.3 ± 0.2 obtained for wild-type cytochrome c′. However, the stabilities of all histidine-heme loops, which contain the helix 3 sequence segment, are increased in the A83S/A87S background compared to the wild-type background. Rate constants for histidine-heme loop breakage are similar for the wild-type and A83S/A87S variants. However, for histidine-heme loops that contain the helix 3 sequence segment, the rate constants for loop formation increase in the A83S/A87S background compared to the wild-type background. Thus, residual helical structure appears to stiffen the polypeptide chain slowing loop formation in the denatured state. The implications of these results for protein folding mechanisms are discussed.  相似文献   

2.
3.
Integrase (IN) is the catalytic component of the preintegration complex, a large nucleoprotein assembly critical for the integration of the retroviral genome into a host chromosome. Although partial crystal structures of human immunodeficiency virus IN alone and its complex with the integrase binding domain of the host factor PSIP1/lens epithelium-derived growth factor (LEDGF)/p75 are available, many questions remain regarding the properties and structures of LEDGF-bound IN oligomers. Using analytical ultracentrifugation, multiangle light scattering, and small angle x-ray scattering, we have established the oligomeric state, stoichiometry, and molecular shapes of IN·LEDGF complexes in solution. Analyses of intact IN tetramers bound to two different LEDGF truncations allow for placement of the integrase binding domain by difference analysis. Modeling of the small angle x-ray scattering envelopes using existing structural data suggests domain arrangements in the IN oligomers that support and extend existing biochemical data for IN·LEDGF complexes and lend new insights into the quaternary structure of LEDGF-bound IN tetramers. These IN oligomers may be involved in stages of the viral life cycle other than integration, including assembly, budding, and early replication.  相似文献   

4.
The efficacy of agonists at Cys-loop ion channel receptors is determined by the rate they isomerize receptors to a pre-open flip state. Once the flip state is reached, the shut-open reaction is similar for low and high efficacy agonists. The present study sought to identify a conformational change associated with the closed-flip transition in the α1-glycine receptor. We employed voltage-clamp fluorometry to compare ligand-binding domain conformational changes induced by the following agonists, listed from highest to lowest affinity and efficacy: glycine > β-alanine > taurine. Voltage-clamp fluorometry involves labeling introduced cysteines with environmentally sensitive fluorophores and inferring structural rearrangements from ligand-induced fluorescence changes. Agonist affinity and efficacy correlated inversely with maximum fluorescence magnitudes at labeled residues in ligand-binding domain loops D and E, suggesting that large conformational changes in this region preclude efficacious gating. However, agonist affinity and efficacy correlated directly with maximum fluorescence magnitudes from a label attached to A52C in loop 2, near the transmembrane domain interface. Because glycine experiences the largest affinity increase between closed and flip states, we propose that the magnitude of this fluorescence signal is directly proportional to the agonist affinity increase. In contrast, labeled residues in loops C, F, and the pre-M1 domain yielded agonist-independent fluorescence responses. Our results support the conclusion that a closed-flip conformation change, with a magnitude proportional to the agonist affinity increase from closed to flip states, occurs in the microenvironment of Ala-52.Glycine receptors (GlyRs)3 are pentameric chloride-selective ion channels that mediate fast inhibitory neurotransmission (1). They are members of the Cys-loop receptor family that includes the prototypical nicotinic acetylcholine receptor (nAChR), the γ-aminobutyric acid type-A receptors (GABAARs), and serotonin type-3 receptors (5-HT3Rs). Recent structural studies have provided a wealth of information on the structure and function of this receptor family (26). In Cys-loop receptors, the ligand-binding domain (LBD) preceding the four transmembrane helices consists of two twisted β-sheets. The inner (vestibule facing) β-sheet comprises seven β-strands, while the outer β-sheet is formed by three β-strands (3). The ligand binding site is located at the interface of adjacent subunits and is lined by six domains: three loops from the principal and the complementary sides, termed A-C and D-F, respectively (3).GlyRs are activated by endogenous amino acid agonists in the following order of efficacy: glycine > β-alanine > taurine (7, 8). As these amino acids share considerable structural similarity (Fig. 1A), they are likely to compete for the same binding site (911). A recent ground-breaking study on an intermediate pre-open state, the so-called “flip” state (12), has provided new insights into the mechanism of partial agonism in Cys-loop receptors (13). This study suggested that agonist efficacy depends on the ability of the agonist to convert the inert agonist-bound receptor to the pre-open flip state. Once the flip state is reached, the shut-open reaction is similar for high and low efficacy agonists. To date there is, however, very little information concerning the structural basis for the lower efficacies of partial agonists. To address this, the present study employed the voltage-clamp fluorometry (VCF) technique (14) to compare the conformational changes induced by glycine, β-alanine, and taurine at various positions in the GlyR LBD.Open in a separate windowFIGURE 1.A, structures of glycine, β-alanine, and taurine. B, model of the LBD, based on carbomylcholine-bound AChBP (PDB code 1uv6). The inner β-sheet is displayed in red, the outer β-sheet in blue. Connecting loops are shown in gray. Colored balls represent approximate locations of selected residues labeled in regions flanking the outer β-sheet (black, G181C in loop F; N203C in loop C; Q219C in the pre-M1 domain) and in the inner β-sheet (yellow, L127C in loop E; Q67C in loop D; A52C in loop 2).VCF involves tethering of an environmentally sensitive fluorophore to a cysteine engineered into a domain of interest. If ligand-binding and/or channel opening leads to a changed dielectric environment surrounding the fluorophore, a change in quantum yield or emission spectrum can be detected. VCF was first employed on voltage-gated potassium channels (15) and has since provided a wealth of information on Cys-loop receptor structure and function (1623). Here we employ VCF to identify an agonist-specific conformational change that may control or reflect the rate at which the GlyR isomerizes to the flip state.  相似文献   

5.
The GABA transporter-1 (GAT-1) has three current-generating modes: GABA-coupled current, Li+-induced leak current, and Na+-dependent transient currents. We earlier hypothesized that Li+ is able to substitute for the first Na+ in the transport cycle and thereby induce a distinct conformation in GAT-1 and that the onset of the Li+-induced leak current at membrane potentials more negative than −50 mV was due to a voltage-dependent conformational change of the Li+-bound transporter. In this study, we set out to verify this hypothesis and seek insight into the structural dynamics underlying the leak current, as well as the sodium-dependent transient currents, by applying voltage clamp fluorometry to tetramethylrhodamine 6-maleimide-labeled GAT-1 expressed in Xenopus laevis oocytes. MTSET accessibility studies demonstrated the presence of two distinct conformations of GAT-1 in the presence of Na+ or Li+. The voltage-dependent fluorescence intensity changes obtained in Li+ buffer correlated with the Li+-induced leak currents, i.e. both were highly voltage-dependent and only present at hyperpolarized potentials (<−50 mV). The transient currents correlated directly with the voltage-dependent fluorescence data obtained in sodium buffer and the associated conformational changes were distinct from those associated with the Li+-induced leak current. The inhibitor potency of SKF89976A of the Li+- versus Na+-bound transporter confirmed the cationic dependence of the conformational occupancy. Our observations suggest that the microdomain situated at the external end of transmembrane I is involved in different conformational changes taking place either during the binding and release of sodium or during the initiation of the Li+-induced leak current.γ-Aminobutyric acid (GABA)2 is the major inhibitory neurotransmitter in the mammalian central nervous system. Continuous GABAergic neurotransmission is efficiently prevented by a GABA re-uptake system that transports GABA back into the synaptic processes via the GABA transporters (GAT). Four isoforms of the mammalian GAT have been found: GAT-1, GAT-2, GAT-3, and BGT-1 (betaine transporter-1) (1). These membrane proteins couple the transport of one GABA molecule to the transport of two Na+ and one Cl (2, 3). Accordingly, the transport process is electrogenic and the transport activity can therefore be monitored by electrophysiological methods. GAT-1 has also been shown to generate: (i) an inwardly rectifying leak current in the presence of Li+ (and in complete absence of Na+) when the membrane potential is more negative than −50 mV (46) and (ii) a presteady-state transient current in the presence of Na+ but in the absence of GABA in response to step jumps in membrane voltage (4, 7).GAT-1 is strictly dependent on external Na+ to drive the transport of GABA (7) and external GABA does not affect the Li+-induced leak current (8). Taken together, this suggests that Li+ cannot induce the same conformation in GAT-1 as Na+ is capable of inducing: namely the conformation that is required for the binding and translocation of GABA. This is in contrast to, e.g. the related serotonin transporter and dopamine transporter in which substrate inhibits the Li+-induced leak current (9, 10) and the Na+/glucose cotransporter (SGLT1) where Li+ is able to sustain substrate transport (11). We have in an earlier study shown that Li+ is able to bind to the first low apparent affinity cation-binding site in the transport cycle (and replace Na+) but not to the second high apparent affinity cation-binding site (8). This finding was later confirmed by Kanner and co-workers (12), who were able to identify the cation-binding site with which Li+ interacts and is able to replace Na+. According to our model, the transporter in the presence of Li+ is “stalled” in the conformation in which only the first cation-binding site is occupied, in contrast to the presence of Na+, where both cation-binding sites are occupied.An unresolved question on the Li+-induced leak current for GAT-1 is the mechanism of the prominent inward rectification. The electrochemical driving force for Li+ predicts a Li+-induced inward current originating at much more positive membrane potentials, but this current is not detected unless the membrane potential is more negative than −50 mV. We previously hypothesized that the Li+-induced leak mode would commence only at the hyperpolarized membrane potentials due to a voltage-dependent conformational change in the Li+-bound GABA transporter (8). In the present study, we set out to test this hypothesis by introduction of a fluorescent probe in GAT-1 to monitor voltage-dependent local conformational changes and relate these to the different current-generating modes of GAT-1.We expressed GAT-1 in Xenopus laevis oocytes and used simultaneous electrical and optical measurements (voltage clamp fluorometry) (13) to monitor the currents and conformational changes of the transporter in the presence of Li+ and Na+ in response to step changes in membrane potential. By labeling Cys74 at the external end of transmembrane helix (TM) I of rat GAT-1 (see Fig. 1) with the cysteine-reactive fluorescent probe, tetramethylrhodamine 6-maleimide (TMR6M), we were able to correlate the voltage dependence of the Li+-induced leak current and the Li+- and voltage-dependent changes in conformations observed by fluorescence intensity changes. The voltage dependence of the Li+-induced conformational changes appeared distinct from the Na+-induced conformational changes associated with the Na+-dependent transient currents. We also explored differences in the inhibitor potency of the GAT-1-specific inhibitor SKF89976A (14) as well as the differential inhibition of the GABA transport by the cysteine-reactive methanethiosulfonate ethyltrimethylammonium (MTSET) in the presence of either Na+ or Li+. Finally, we prepared a homology model of GAT-1 (Fig. 1) by using the bacterial leucine transporter, LeuTAa (15), as a template and dock the TMR6M into the model to provide a framework for interpreting the putative conformational rearrangements that may explain the observed changes in fluorescence intensity.Open in a separate windowFIGURE 1.Three-dimensional model of GAT-1 with the fluorophore, TMR6M, covalently attached to Cys74. The model was made by homology modeling with the bacterial LeuTAa transporter as template. A, side view of the GAT-1 model. The 12 transmembrane helices are shown in different colors; TM1 being blue and TM12 being red. The two sodium ions are purple spheres, chloride is a green sphere, and GABA is shown next to the sodium ions as red, light and dark blue spheres. TMR6M is located in the external surface of the model (shown as green, blue, and red spheres) and is attached to Cys74 (shown as blue, red, and yellow spheres). B, magnified view of the local environment of TMR6M embedded in a hydrophobic cleft between EL3 (green), the beginning of EL4 (yellow), and the outer part of TM1 (blue). Below TMR6M, sodium, chloride, and GABA can be seen as spheres.Altogether, the present data support that local conformational changes taking place at the external surface of TM1 to mirror the global conformational changes taking place during the current-generating modes of the GABA transporter. Moreover, our data demonstrate that voltage dependence of the conformational changes associated with the Li+-induced leak current is different from the Na+-dependent conformational changes required for GABA transport.  相似文献   

6.
Understanding the activation mechanism of Cys loop ion channel receptors is key to understanding their physiological and pharmacological properties under normal and pathological conditions. The ligand-binding domains of these receptors comprise inner and outer β-sheets and structural studies indicate that channel opening is accompanied by conformational rearrangements in both β-sheets. In an attempt to resolve ligand-dependent movements in the ligand-binding domain, we employed voltage-clamp fluorometry on α1 glycine receptors to compare changes mediated by the agonist, glycine, and by the antagonist, strychnine. Voltage-clamp fluorometry involves labeling introduced cysteines with environmentally sensitive fluorophores and inferring structural rearrangements from ligand-induced fluorescence changes. In the inner β-sheet, we labeled residues in loop 2 and in binding domain loops D and E. At each position, strychnine and glycine induced distinct maximal fluorescence responses. The pre-M1 domain responded similarly; at each of four labeled positions glycine produced a strong fluorescence signal, whereas strychnine did not. This suggests that glycine induces conformational changes in the inner β-sheet and pre-M1 domain that may be important for activation, desensitization, or both. In contrast, most labeled residues in loops C and F yielded fluorescence changes identical in magnitude for glycine and strychnine. A notable exception was H201C in loop C. This labeled residue responded differently to glycine and strychnine, thus underlining the importance of loop C in ligand discrimination. These results provide an important step toward mapping the domains crucial for ligand discrimination in the ligand-binding domain of glycine receptors and possibly other Cys loop receptors.Glycine receptor (GlyR)3 chloride channels are pentameric Cys loop receptors that mediate fast synaptic transmission in the nervous system (1, 2). This family also includes nicotinic acetylcholine receptors (nAChRs), γ-aminobutyric acid type A and type C receptors, and serotonin type 3 receptors. Individual subunits comprise a large ligand-binding domain (LBD) and a transmembrane domain consisting of four α-helices (M1–M4). The LBD consists of a 10-strand β-sandwich made of an inner β-sheet with six strands and an outer β-sheet with four strands (3). The ligand-binding site is situated at the interface of adjacent subunits and is formed by loops A–C from one subunit and loops D–F from the neighboring subunit (3).The activation mechanism of Cys loop receptors is currently the subject of intense investigation because it is key to understanding receptor function under normal and pathological conditions (4, 5). Based on structural analysis of Torpedo nAChRs, Unwin and colleagues (6, 7) originally proposed that agonist binding induced the inner β-sheet to rotate, whereas the outer β-sheet tilted slightly upwards with loop C clasping around the agonist. These movements were thought to be transmitted to the transmembrane domain via a differential movement of loop 2 (β1-β2) and loop 7 (β6-β7) (both part of the inner β-sheet) and the pre-M1 domain (which is linked via a β-strand to the loop C in the outer sheet). The idea of large loop C movements accompanying agonist binding is supported by structural and functional data (3, 813). However, a direct link between loop C movements and channel gating has proved more difficult to establish. Although computational modeling studies have suggested that this loop may be a major component of the channel opening mechanism (1418), experimental support for this model is not definitive. Similarly, loop F is also thought to move upon ligand binding, although there is as yet no consensus as to whether these changes represent local or global conformational changes (11, 1921). Recently, a comparison of crystal structures of bacterial Cys loop receptors in the closed and open states revealed that although both the inner and outer β-sheets exhibit different conformations in closed and open states, the pre-M1 domain remains virtually stationary (22, 23). It is therefore relevant to question whether loop C, loop F, and pre-M1 movements are essential for Cys loop receptor activation.Strychnine is a classical competitive antagonist of GlyRs (24, 25), and to date there is no evidence that it can produce LBD structural changes. In this study we use voltage-clamp fluorometry (VCF) to compare glycine- and strychnine-induced conformational changes in the GlyR loops 2, C, D, E, and F and the pre-M1 domain in an attempt to determine whether they signal ligand-binding events, local conformational changes, or conformational changes associated with receptor activation.In a typical VCF experiment, a domain of interest is labeled with an environmentally sensitive fluorophore, and current and fluorescence are monitored simultaneously during ligand application. VCF is ideally suited for identifying ligand-specific conformational changes because it can report on electrophysiologically silent conformational changes (26), such as those induced by antagonists. Indeed, VCF has recently provided valuable insights into the conformational rearrangements of various Cys loop receptors (19, 21, 2733).  相似文献   

7.
8.
We used a combination of fluorescence, circular dichroism (CD), and NMR spectroscopies in conjunction with size exclusion chromatography to help rationalize the relative antibacterial, antiplasmodial, and cytotoxic activities of a series of proline-free and proline-containing model antimicrobial peptides (AMPs) in terms of their structural properties. When compared with proline-free analogs, proline-containing peptides had greater activity against Gram-negative bacteria, two mammalian cancer cell lines, and intraerythrocytic Plasmodium falciparum, which they were capable of killing without causing hemolysis. In contrast, incorporation of proline did not have a consistent effect on peptide activity against Mycobacterium tuberculosis. In membrane-mimicking environments, structures with high α-helix content were adopted by both proline-free and proline-containing peptides. In solution, AMPs generally adopted disordered structures unless their sequences comprised more hydrophobic amino acids or until coordinating phosphate ions were added. Proline-containing peptides resisted ordering induced by either method. The roles of the angle subtended by positively charged amino acids and the positioning of the proline residues were also investigated. Careful positioning of proline residues in AMP sequences is required to enable the peptide to resist ordering and maintain optimal antibacterial activity, whereas varying the angle subtended by positively charged amino acids can attenuate hemolytic potential albeit with a modest reduction in potency. Maintaining conformational flexibility improves AMP potency and selectivity toward bacterial, plasmodial, and cancerous cells while enabling the targeting of intracellular pathogens.  相似文献   

9.
In Escherichia coli, the γ complex clamp loader loads the β-sliding clamp onto DNA. The β clamp tethers DNA polymerase III to DNA and enhances the efficiency of replication by increasing the processivity of DNA synthesis. In the presence of ATP, γ complex binds β and DNA to form a ternary complex. Binding to primed template DNA triggers γ complex to hydrolyze ATP and release the clamp onto DNA. Here, we investigated the kinetics of forming a ternary complex by measuring rates of γ complex binding β and DNA. A fluorescence intensity-based β binding assay was developed in which the fluorescence of pyrene covalently attached to β increases when bound by γ complex. Using this assay, an association rate constant of 2.3 × 107 m−1 s−1 for γ complex binding β was determined. The rate of β binding was the same in experiments in which γ complex was preincubated with ATP before adding β or added directly to β and ATP. In contrast, when γ complex is preincubated with ATP, DNA binding is faster than when γ complex is added to DNA and ATP at the same time. Slow DNA binding in the absence of ATP preincubation is the result of a rate-limiting ATP-induced conformational change. Our results strongly suggest that the ATP-induced conformational changes that promote β binding and DNA binding differ. The slow ATP-induced conformational change that precedes DNA binding may provide a kinetic preference for γ complex to bind β before DNA during the clamp loading reaction cycle.  相似文献   

10.
Conformational changes of αs-casein by heating were investigated by measuring ultraviolet difference spectra. The ultraviolet difference spectra at elevated temperature against 5.5°C were measured in various ionic strengths and pHs. Thermal effects of the difference spectra were cancelled by comparing with the spectra of model compounds such as lysozyme and ribonuclease, and the blue shift of αs-casein spectra was observed at above 30°C in these all experimental conditions. This shift was considered to mean unfolding of the αs-casein molecule. The aggregation of αs-casein was observed above ionic strength of 0.4 by heating. These heat-induced changes were reversible until the aggregation was observed.  相似文献   

11.
The transient protein-protein interactions induced by guanine nucleotide-dependent conformational changes of G proteins play central roles in G protein-coupled receptor-mediated signaling systems. Leukemia-associated RhoGEF (LARG), a guanine nucleotide exchange factor for Rho, contains an RGS homology (RH) domain and Dbl homology/pleckstrin homology (DH/PH) domains and acts both as a GTPase-activating protein (GAP) and an effector for Gα13. However, the molecular mechanism of LARG activation upon Gα13 binding is not yet well understood. In this study, we analyzed the Gα13-LARG interaction using cellular and biochemical methods, including a surface plasmon resonance (SPR) analysis. The results obtained using various LARG fragments demonstrated that active Gα13 interacts with LARG through the RH domain, DH/PH domains, and C-terminal region. However, an alanine substitution at the RH domain contact position in Gα13 resulted in a large decrease in affinity. Thermodynamic analysis revealed that binding of Gα13 proceeds with a large negative heat capacity change (ΔCp°), accompanied by a positive entropy change (ΔS°). These results likely indicate that the binding of Gα13 with the RH domain triggers conformational rearrangements between Gα13 and LARG burying an exposed hydrophobic surface to create a large complementary interface, which facilitates complex formation through both GAP and effector interfaces, and activates the RhoGEF. We propose that LARG activation is regulated by an induced-fit mechanism through the GAP interface of Gα13.Heterotrimeric G proteins3 serve as key molecular switches to transduce a large array of extracellular signals into cells by actively alternating their conformations between GDP-bound inactive and GTP-bound active forms. In the current model, the ligand-activated G protein-coupled receptors (GPCRs) catalyze the exchange of GDP for GTP on Gα subunits (1). Upon activation, three switch regions in the Gα subunit undergo significant conformational changes, followed by dissociation of the GTP-bound Gα subunit from the Gβγ subunits. Both Gα-GTP and free Gβγ interact with diverse downstream effectors to transmit intracellular signals. The Gα subunit hydrolyzes bound GTP to GDP by its intrinsic GTPase activity. This deactivation process is further accelerated by GTPase-activating proteins (GAPs) such as regulator of G protein signaling (RGS) proteins (2, 3). Gα-GDP dissociates from effectors and re-associates with Gβγ to terminate the signal.Although this model explains the basic concept of G protein signaling, the molecular dynamics of interactions among GPCR, G protein, RGS protein, and effector during the signaling process is not well understood. It has been suggested that the GPCR signals are integrated into the intracellular signaling network at the level of G proteins (4). Accumulating evidence suggests that the Gα subunit acts as the core of the signaling complex at the membrane, which is formed through the transient protein-protein interactions of multiple signaling components (5, 6). Thus, the quantitative analysis of the dynamic molecular interactions in the GPCR signaling complex will be crucial to understanding various cellular processes.Gα12 and Gα13 subunits have been demonstrated to regulate the activity of Rho GTPase through RhoGEFs, which contain an N-terminal RGS homology domain (RH-RhoGEFs) (710). RH-RhoGEFs, which consist of p115RhoGEF/Lsc, PDZ-Rho-GEF/GTRAP48, and LARG in mammalian species, directly link the activation of GPCRs by extracellular ligands to the regulation of Rho activity in cells (1014). All three RH-RhoGEFs contain an N-terminal RH domain, which specifically recognizes the active form of Gα12 or Gα13 and central DH/PH domains characteristic of GEFs for Rho GTPases. It has been demonstrated in vitro that LARG and p115RhoGEF serve as specific GAPs for Gα12/13 through their RH domains and also as their effectors to regulate Rho GTPase activation (1113). A structural study has demonstrated that the interface of the RH domain of p115RhoGEFs and a Gα13/i1 chimera is different from that of the RGS domain of RGS4 and Gαi1 (7). The N-terminal small element in the RH domain, which is required for GAP activity toward Gα13, contacts the switch regions and the helical domain of the Gα13/i1 chimera. The core module of the p115RhoGEF RH domain binds to the region of Gα13/i1, which is conventionally used for effector binding. These results suggest roles for the RH domain in the stimulation of GEF activity by Gα13 in addition to GAP activity. On the other hand, several studies have also indicated that regions outside of RH domain of RH-RhoGEFs, particularly the DH/PH domains, interact directly with activated Gα13 (11, 14, 15). In addition, we have demonstrated recently that p115RhoGEF interacts with distinct surfaces of Gα13 for the GAP reaction or GEF activity regulation (16). However, the molecular mechanism of LARG activation upon Gα13 binding is not clearly understood.In this study, we have developed a quantitative method for the kinetic and thermodynamic analysis of Gα13-effector interaction using surface plasmon resonance (SPR) with sensor chips on which Gα13 was immobilized. We examined the kinetics and thermodynamics of the Gα13-LARG interaction and assessed LARG activation using both in vitro and cell-based approaches. We present evidence that, in addition to the interaction with the RH domain, the DH/PH domains and C-terminal region of LARG also interact with Gα13 to form the high affinity Gα13-LARG complex and activate RhoGEF activity. We further propose that LARG adopts the active conformation using an induced-fit mechanism through association with the GAP interface of Gα13. A similar mechanism may also be used with other Gα-effector interactions.  相似文献   

12.
13.
Bridged bicyclic amino acids have high potential applicability as self-organized, conformationally constrained synthetic building blocks that do not require assistance from hydrogen bond formation. We systematically investigated the intrinsic conformational propensities of dipeptides of bridged bicyclic β-amino acids by means of accelerated molecular dynamics simulation and density functional theory (DFT) calculations in methanol, chloroform, and water. While the main-chain conformation, represented by φ and θ values, is fixed by the nature of the bicyclic ring structure, rotation of the C-terminal carbonyl group (ψ) is also restricted, converging to one or two minima. In endo-type dipeptides, in which the two N- and C-terminal amides are spatially close to each other, the C-terminal amide plane is placed horizontally. In exo-type dipeptides, in which the two amides are on opposite sides of the ring plane, the C-terminal carbonyl group can take two types of positions: either parallel/antiparallel with the N-terminal carbonyl or beneath the bicyclic ring, forcing the amide NHMe moiety to lie outside of the ring. We also examined the cis-trans preference of model bicyclic amides. Although the parent amides exhibit cis-trans equilibrium without any preference, addition of a methyl group on one of the bridgehead positions tips the equilibrium towards trans.  相似文献   

14.
The angiotensin-converting enzyme (ACE) is a zinc-dependent metalloproteinase widely occurring in the organism; it metabolizes many peptides and plays a key role in blood pressure regulation and vascular remodeling. This enzyme is expressed as a type-1 membrane glycoprotein on the surface of endothelial and epithelial cells, but is also found in a soluble form in biological fluids. In this study, we used purified ACE from lungs, which is mainly produced by endothelial cells of lung capillaries; ACE from heart, produced by endothelial heart cells and, probably, by myofibroblasts; and ACE from seminal fluid, produced by the epithelial cells of the prostate and epididymis. The pattern of binding of a set of 17 mAbs to different conformational epitopes on the surface of two domains of the human ACE significantly differed for ACEs from different organs. This pattern (the conformational “fingerprint” of ACE) reflects the local conformation of the surface of a particular ACE. The differences in the conformational fingerprints of ACEs expressed by different cell types, or even by similar cells but in different organs, can be explained by the posttranslational modification of ACE protein in these organs and, primarily, different glycosylation of N-glycosylation sites Asn25, Asn117, Asn289, Asn666, Asn685, and Asn731. The mass spectrometry of tryptic hydrolyzates of ACEs isolated from different human organs made it possible to reveal, in the composition of different ACEs, N-glycosylation sites that are really occupied by glycans, namely, Asn in positions 82, 117, 416, 648, 666, 685, and 731 in ACE from seminal fluid; Asn in positions 117, 648, 666, and 685 in ACE from lungs; and Asn in positions 117, 480, 666, and 685 in ACE from heart. Differences in the plausible structures of glycans in ACE, in particular, at the Asn666 N-glycosylation site were demonstrated, which can explain the differences in the efficiency of binding of mAbs to ACE from different organs.  相似文献   

15.
The conformation of a melittin—inhibitor complex was studied by solution NMR, solid-state NMR, and circular dichroism. In solution, binding was studied by titrating inhibitor against melittin in dimethyl sulfoxide, methanol, aqueous buffer, and dodecylphosphocholine micelles. The change in chemical shift of Trp19 resonances and the formation of a precipitate at 1:1 molar ratio indicated that the inhibitor was bound to melittin. Solid-state NMR also showed a change in chemical shift of two labeled carbons of melittin near Pro14 and a change in 1H T 1 relaxation times when complexed with inhibitor. Rotational resonance experiments of melittin labeled in the proline region indicated a change in conformation for melittin complexed with inhibitor. This observation was also supported by circular dichroism measurements, indicating a reduction in α-helical structure for increasing ratios of inhibitor bound to melittin.  相似文献   

16.
17.
Synapses are asymmetric intercellular junctions that mediate neuronal communication. The number, type, and connectivity patterns of synapses determine the formation, maintenance, and function of neural circuitries. The complexity and specificity of synaptogenesis relies upon modulation of adhesive properties, which regulate contact initiation, synapse formation, maturation, and functional plasticity. Disruption of adhesion may result in structural and functional imbalance that may lead to neurodevelopmental diseases, such as autism, or neurodegeneration, such as Alzheimer''s disease. Therefore, understanding the roles of different adhesion protein families in synapse formation is crucial for unraveling the biology of neuronal circuit formation, as well as the pathogenesis of some brain disorders. The present review summarizes some of the knowledge that has been acquired in vertebrate and invertebrate genetic model organisms.Synapses are asymmetric, intercellular junctions that are the basic structural units of neuronal transmission. The correct development of synaptic specializations and the establishment of appropriate connectivity patterns are crucial for the assembly of functional neuronal circuits. Improper synapse formation and function may cause neurodevelopmental disorders, such as mental retardation (MsR) and autism spectrum disorders (ASD) (McAllister 2007; Sudhof 2008), and likely play a role in neurodegenerative disorders, such as Alzheimer''s disease (AD) (Haass and Selkoe 2007).At chemical synapses (reviewed in Sudhof 2004; Zhai and Bellen 2004; Waites et al. 2005; McAllister 2007; Jin and Garner 2008), the presynaptic compartment contains synaptic vesicles (SV), organized in functionally distinct subcellular pools. A subset of SVs docks to the presynaptic membrane around protein-dense release sites, named active zones (AZ). Upon the arrival of an action potential at the terminal, the docked and “primed” SVs fuse with the plasma membrane and release neurotransmitter molecules into the synaptic cleft. Depending on the type of synapse (i.e., excitatory vs. inhibitory synapses), neurotransmitters ultimately activate an appropriate set of postsynaptic receptors that are accurately apposed to the AZ.Synapse formation occurs in several steps (Fig. 1) (reviewed in Eaton and Davis 2003; Goda and Davis 2003; Waites et al. 2005; Garner et al. 2006; Gerrow and El-Husseini 2006; McAllister 2007). Spatiotemporal signals guide axons through heterogeneous cellular environments to contact appropriate postsynaptic targets. At their destination, axonal growth cones initiate synaptogenesis through adhesive interactions with target cells. In the mammalian central nervous system (CNS), immature postsynaptic dendritic spines initially protrude as thin, actin-rich filopodia on the surface of dendrites. Similarly, at the Drosophila neuromuscular junction (NMJ), myopodia develop from the muscles (Ritzenthaler et al. 2000). The stabilization of intercellular contacts and their elaboration into mature, functional synapses involves cytoskeletal arrangements and recruitment of pre- and postsynaptic components to contact sites in spines and boutons. Conversely, retraction of contacts results in synaptic elimination. Both stabilization and retraction sculpt a functional neuronal circuitry.Open in a separate windowFigure 1.(A–C) Different stages of synapse formation. (A) Target selection, (B) Synapse assembly, (C) Synapse maturation and stabilization. (D–F) The role of cell adhesion molecules in synapse formation is exemplified by the paradigm of N-cadherin and catenins in regulation of the morphology and strength of dendritic spine heads. (D) At an early stage the dendritic spines are elongated from motile structures “seeking” their synaptic partners. (E) The contacts between the presynaptic and postsynaptic compartments are stabilized by recruitment of additional cell adhesion molecules. Adhesional interactions activate downstream pathways that remodel the cytoskeleton and organize pre- and postsynaptic apparatuses. (F) Cell adhesion complexes, stabilized by increased synaptic activity, promote the expansion of the dendritic spine head and the maturation/ stabilization of the synapse. Retraction and expansion is dependent on synaptic plasticity.In addition to the plastic nature of synapse formation, the vast heterogeneity of synapses (in terms of target selection, morphology, and type of neurotransmitter released) greatly enhances the complexity of synaptogenesis (reviewed in Craig and Boudin 2001; Craig et al. 2006; Gerrow and El-Husseini 2006). The complexity and specificity of synaptogenesis relies upon the modulation of adhesion between the pre- and postsynaptic components (reviewed in Craig et al. 2006; Gerrow and El-Husseini 2006; Piechotta et al. 2006; Dalva et al. 2007; Shapiro et al. 2007; Yamada and Nelson 2007; Gottmann 2008). Cell adhesive interactions enable cell–cell recognition via extracellular domains and also mediate intracellular signaling cascades that affect synapse morphology and organize scaffolding complexes. Thus, cell adhesion molecules (CAMs) coordinate multiple synaptogenic steps.However, in vitro and in vivo studies of vertebrate CAMs are often at odds with each other. Indeed, there are no examples of mutants for synaptic CAMs that exhibit prominent defects in synapse formation. This apparent “resilience” of synapses is probably caused by functional redundancy or compensatory effects among different CAMs (Piechotta et al. 2006). Hence, studies using simpler organisms less riddled by redundancy, such as Caenorhabditis elegans and Drosophila, have aided in our understanding of the role that these molecules play in organizing synapses.In this survey, we discuss the roles of the best characterized CAM families of proteins involved in synaptogenesis. Our focus is to highlight the complex principles that govern the molecular basis of synapse formation and function from a comparative perspective. We will present results from cell culture studies as well as in vivo analyses in vertebrate systems and refer to invertebrate studies, mainly performed in Drosophila and C. elegans, when they have provided important insights into the role of particular CAM protein families. However, we do not discuss secreted factors, for which we refer the reader to numerous excellent reviews (as for example Washbourne et al. 2004; Salinas 2005; Piechotta et al. 2006; Shapiro et al. 2006; Dalva 2007; Yamada and Nelson 2007; Biederer and Stagi 2008; Salinas and Zou 2008).  相似文献   

18.
Alzheimer disease (AD) is a devastating neurodegenerative disease with complex and strong genetic inheritance. Four genes have been established to either cause familial early onset AD (APP, PSEN1, and PSEN2) or to increase susceptibility for late onset AD (APOE). To date ∼80% of the late onset AD genetic variance remains elusive. Recently our genome-wide association screen identified four novel late onset AD candidate genes. Ataxin 1 (ATXN1) is one of these four AD candidate genes and has been indicated to be the disease gene for spinocerebellar ataxia type 1, which is also a neurodegenerative disease. Mounting evidence suggests that the excessive accumulation of Aβ, the proteolytic product of β-amyloid precursor protein (APP), is the primary AD pathological event. In this study, we ask whether ATXN1 may lead to AD pathogenesis by affecting Aβ and APP processing utilizing RNA interference in a human neuronal cell model and mouse primary cortical neurons. We show that knock-down of ATXN1 significantly increases the levels of both Aβ40 and Aβ42. This effect could be rescued with concurrent overexpression of ATXN1. Moreover, overexpression of ATXN1 decreased Aβ levels. Regarding the underlying molecular mechanism, we show that the effect of ATXN1 expression on Aβ levels is modulated via β-secretase cleavage of APP. Taken together, ATXN1 functions as a genetic risk modifier that contributes to AD pathogenesis through a loss-of-function mechanism by regulating β-secretase cleavage of APP and Aβ levels.  相似文献   

19.
Protein misfolding due to missense mutations is a common pathogenic mechanism in cystathionine β-synthase (CBS) deficiency. In our previous studies, we successfully expressed, purified, and characterized nine CBS mutant enzymes containing the following patient mutations: P49L, P78R, A114V, R125Q, E176K, R266K, P422L, I435T, and S466L. These purified mutants exhibited full heme saturation, normal tetrameric assembly, and high catalytic activity. In this work, we used several spectroscopic and proteolytic techniques to provide a more thorough insight into the conformation of these mutant enzymes. Far-UV circular dichroism, fluorescence, and second-derivative UV spectroscopy revealed that the spatial arrangement of these CBS mutants is similar to that of the wild type, although the microenvironment of the chromophores may be slightly altered. Using proteolysis with thermolysin under native conditions, we found that the majority of the studied mutants is more susceptible to cleavage, suggesting their increased local flexibility or propensity for local unfolding. Interestingly, the presence of the CBS allosteric activator, S-adenosylmethionine (AdoMet), increased the rate of cleavage of the wild type and the AdoMet-responsive mutants, while the proteolytic rate of the AdoMet-unresponsive mutants was not significantly changed. Pulse proteolysis analysis suggested that the protein structure of the R125Q and E176K mutants is significantly less stable than that of the wild type and the other mutants. Taken together, the proteolytic data shows that the conformation of the pathogenic mutants is altered despite retained catalytic activity and normal tetrameric assembly. This study demonstrates that the proteolytic techniques are useful tools for the assessment of the biochemical penalty of missense mutations in CBS.  相似文献   

20.
A typical plasmid replicon of Escherichia coli, such as ori γ of R6K, contains tandem iterons (iterated initiator protein binding sites), an AT-rich region that melts upon initiator-iteron interaction, two binding sites for the bacterial initiator protein DnaA, and a binding site for the DNA-bending protein IHF. R6K also contains two structurally atypical origins called α and β that are located on either side of γ and contain a single and a half-iteron, respectively. Individually, these sites do not bind to initiator protein π but access it by DNA looping-mediated interaction with the seven π-bound γ iterons. The π protein exists in 2 interconvertible forms: inert dimers and active monomers. Initiator dimers generally function as negative regulators of replication by promoting iteron pairing (“handcuffing”) between pairs of replicons that turn off both origins. Contrary to this existing paradigm, here we show that both the dimeric and the monomeric π are necessary for ori α-driven plasmid maintenance. Furthermore, efficient looping interaction between α and γ or between 2 γ iterons in vitro also required both forms of π. Why does α-γ iteron pairing promote α activation rather than repression? We show that a weak, transitory α-γ interaction at the iteron pairs was essential for α-driven plasmid maintenance. Swapping the α iteron with one of γ without changing the original sequence context that caused enhanced looping in vitro caused a significant inhibition of α-mediated plasmid maintenance. Therefore, the affinity of α iteron for π-bound γ and not the sequence context determined whether the origin was activated or repressed.  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