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.
Fluorescence correlation spectroscopy (FCS) was used to investigate the hydrodynamic and photophysical properties of PR1 (phytofluor red 1), an intensely red fluorescent biliprotein variant of the truncated cyanobacterial phytochrome 1 (Cph1Delta, which consists of the N-terminal 514 amino acids). Single-molecule diffusion measurements showed that PR1 has excellent fluorescence properties at the single-molecule level, making it an interesting candidate for red fluorescent protein fusions. FCS measurements for probing dimer formation in solution over a range of protein concentrations were enabled by addition of Cph1Delta apoprotein (apoCph1Delta) to nanomolar solutions of PR1. FCS brightness analysis showed that heterodimerization of PR1 with apoCph1Delta altered the chemical environment of the PR1 chromophore to further enhance its fluorescence emission. Fluorescence correlation measurements also revealed interactions between apoCph1Delta and the red fluorescent dyes Cy5.18 and Atto 655 but not Alexa Fluor 660. The concentration dependence of protein:dye complex formation indicated that Atto 655 interacted with, or influenced the formation of, the apoCph1 dimer. These studies presage the utility of phytofluor tags for probing single-molecule dynamics in living cells in which the fluorescence signal can be controlled by the addition of various chromophores that have different structures and photophysical properties, thereby imparting different types of information, such as dimer formation or the presence of open binding faces on a protein.  相似文献   

2.
Mechanisms that regulate the nitric oxide synthase enzymes (NOS) are of interest in biology and medicine. Although NOS catalysis relies on domain motions, and is activated by calmodulin binding, the relationships are unclear. We used single-molecule fluorescence resonance energy transfer (FRET) spectroscopy to elucidate the conformational states distribution and associated conformational fluctuation dynamics of the two electron transfer domains in a FRET dye-labeled neuronal NOS reductase domain, and to understand how calmodulin affects the dynamics to regulate catalysis. We found that calmodulin alters NOS conformational behaviors in several ways: It changes the distance distribution between the NOS domains, shortens the lifetimes of the individual conformational states, and instills conformational discipline by greatly narrowing the distributions of the conformational states and fluctuation rates. This information was specifically obtainable only by single-molecule spectroscopic measurements, and reveals how calmodulin promotes catalysis by shaping the physical and temporal conformational behaviors of NOS.Although proteins adopt structures determined by their amino acid sequences, they are not static objects and fluctuate among ensembles of conformations (1). Transitions between these states can occur on a variety of length scales (Å to nm) and time scales (ps to s) and have been linked to functionally relevant phenomena such as allosteric signaling, enzyme catalysis, and protein–protein interactions (24). Indeed, protein conformational fluctuations and dynamics, often associated with static and dynamic inhomogeneity, are thought to play a crucial role in biomolecular functions (511). It is difficult to characterize such spatially and temporally inhomogeneous dynamics in bulk solution by an ensemble-averaged measurement, especially in proteins that undergo multiple-conformation transformations. In such cases, single-molecule spectroscopy is a powerful approach to analyze protein conformational states and dynamics under physiological conditions, and can provide a molecular-level perspective on how a protein’s structural dynamics link to its functional mechanisms (1221).A case in point is the nitric oxide synthase (NOS) enzymes (2224), whose nitric oxide (NO) biosynthesis involves electron transfer reactions that are associated with relatively large-scale movement (tens of Å) of the enzyme domains (Fig. 1A). During catalysis, NADPH-derived electrons first transfer into an FAD domain and an FMN domain in NOS that together comprise the NOS reductase domain (NOSr), and then transfer from the FMN domain to a heme group that is bound in a separate attached “oxygenase” domain, which then enables NO synthesis to begin (22, 2527). The electron transfers into and out of the FMN domain are the key steps for catalysis, and they appear to rely on the FMN domain cycling between electron-accepting and electron-donating conformational states (28, 29) (Fig. 1B). In this model, the FMN domain is suggested to be highly dynamic and flexible due to a connecting hinge that allows it to alternate between its electron-accepting (FAD→FMN) or closed conformation and electron-donating (FMN→heme) or open conformation (Fig. 1 A and B) (28, 3036). In the electron-accepting closed conformation, the FMN domain interacts with the NADPH/FAD domain (FNR domain) to receive electrons, whereas in the electron donating open conformation the FMN domain has moved away to expose the bound FMN cofactor so that it may transfer electrons to a protein acceptor like the NOS oxygenase domain, or to a generic protein acceptor like cytochrome c. In this way, the reductase domain structure cycles between closed and open conformations to deliver electrons, according to a conformational equilibrium that determines the movements and thus the electron flux capacity of the FMN domain (25, 28, 32, 34, 35, 37). A similar conformational switching mechanism is thought to enable electron transfer through the FMN domain in the related flavoproteins NADPH-cytochrome P450 reductase and methionine synthase reductase (3842).Open in a separate windowFig. 1.(A) The nNOSr ribbon structure (from PDB: 1TLL) showing bound FAD (yellow) in FNR domain (green), FMN (orange) in FMN domain (yellow), connecting hinge (blue), and the Cy3–Cy5 label positions (pink) and distance (42 Å, dashed line). (B) Cartoon of an equilibrium between the FMN-closed and FMN-open states, with Cy dye label positions indicated. (C) Cytochrome c reductase activity of nNOSr proteins in their CaM-bound and CaM-free states. Color scheme of bar graphs: Black, WT nNOSr unlabeled; Red, Cys-lite (CL) nNOSr unlabeled; Blue, E827C/Q1268C CL nNOSr unlabeled; and Dark cyan, E827C/Q1268C CL nNOSr labeled.NOS enzymes also contain a calmodulin (CaM) binding domain that is located just before the N terminus of the FMN domain (Fig. 1B), and this provides an important layer of regulation (25, 27). CaM binding to NOS enzymes increases electron transfer from NADPH through the reductase domain and also triggers electron transfer from the FMN domain to the NOS heme as is required for NO synthesis (31, 32). The ability of CaM, or similar signaling proteins, to regulate electron transfer reactions in enzymes is unusual, and the mechanism is a topic of interest and intensive study. It has long been known that CaM binding alters NOSr structure such that, on average, it populates a more open conformation (43, 44). Recent equilibrium studies have detected a buildup of between two to four discreet conformational populations in NOS enzymes and in related flavoproteins, and in some cases, have also estimated the distances between the bound FAD and FMN cofactors in the different species (26, 36, 37, 39, 40), and furthermore, have confirmed that CaM shifts the NOS population distribution toward more open conformations (34, 36, 45). Although valuable, such ensemble-averaged results about conformational states cannot explain how electrons transfer through these enzymes, or how CaM increases the electron flux in NOS, because answering these questions requires a coordinate understanding of the dynamics of the conformational fluctuations. Indeed, computer modeling has indicated that a shift toward more open conformations as is induced by CaM binding to nNOS should, on its own, actually diminish electron flux through nNOS and through certain related flavoproteins (38). Despite its importance, measuring enzyme conformational fluctuation dynamics is highly challenging, and as far as we know, there have been no direct measures on the NOS enzymes or on related flavoproteins, nor studies on how CaM binding might influence the conformational fluctuation dynamics in NOS.To address this gap, we used single-molecule fluorescence energy resonance transfer (FRET) spectroscopy to characterize individual molecules of nNOSr that had been labeled at two specific positions with Cyanine 3 (Cy3) donor and Cyanine 5 (Cy5) acceptor dye molecules, regarding their conformational states distribution and the associated conformational fluctuation dynamics, which in turn enabled us to determine how CaM binding impacts both parameters. This work provides a unique perspective and a novel study of the NOS enzymes and within the broader flavoprotein family, which includes the mammalian enzymes methionine synthase reductase (MSR) and cytochrome P450 reductase (CPR), and reveals how CaM’s control of the conformational behaviors may regulate the electron transfer reactions of NOS catalysis.  相似文献   

3.
A recently developed, real-time spectroscopic technique, burst-integrated fluorescence lifetime (BIFL), is shown to be well suited for monitoring the individual molecular conformational dynamics of a single molecule diffusing through the microscopic, open measurement volume (≈10 fl) of a confocal epi-illuminated set-up. In a highly diluted aqueous solution of 20-mer oligonucleotide strand of DNA duplex labeled with the environment-sensitive fluorescent dye tetramethylrhodamine (TMR), fluorescence bursts indicating traces of individual molecules are registered and further subjected to selective burst analysis. The two-dimensional BIFL data allow the identification and detection of different temporally resolved conformational states. A complementary autocorrelation analysis was performed on the time-dependent fluctuations in fluorescence lifetime and intensity. The consistent results strongly support the hypothesized three-state model of the conformational dynamics of the TMR–DNA duplex with a polar, a nonpolar, and a quenching environment of TMR.  相似文献   

4.
The N- and C-terminal six-helix bundles of lactose permease (LacY) form a large internal cavity open on the cytoplasmic side and closed on the periplasmic side with a single sugar-binding site at the apex of the cavity near the middle of the molecule. During sugar/H(+) symport, an outward-facing cavity is thought to open with closing of the inward-facing cavity so that the sugar-binding site is alternately accessible to either face of the membrane. In this communication, single-molecule fluorescence (F?rster) resonance energy transfer is used to test this model with wild-type LacY and a conformationally restricted mutant. Pairs of Cys residues at the ends of two helices on the cytoplasmic or periplasmic sides of wild-type LacY and the mutant were labeled with appropriate donor and acceptor fluorophores, single-molecule fluorescence resonance energy transfer was determined in the absence and presence of sugar, and distance changes were calculated. With wild-type LacY, binding of a galactopyranoside, but not a glucopyranoside, results in a decrease in distance on the cytoplasmic side and an increase in distance on the periplasmic side. In contrast, with the mutant, a more pronounced decrease in distance and in distance distribution is observed on the cytoplasmic side, but there is no change on the periplasmic side. The results are consistent with the alternating access model and indicate that the defect in the mutant is due to impaired ligand-induced flexibility on the periplasmic side.  相似文献   

5.
The ever-improving time and space resolution and molecular detection sensitivity of fluorescence microscopy offer unique opportunities to deepen our insights into the function of chemical and biological catalysts. Because single-molecule microscopy allows for counting the turnover events one by one, one can map the distribution of the catalytic activities of different sites in solid heterogeneous catalysts, or one can study time-dependent activity fluctuations of individual sites in enzymes or chemical catalysts. By experimentally monitoring individuals rather than populations, the origin of complex behavior, e.g., in kinetics or in deactivation processes, can be successfully elucidated. Recent progress of temporal and spatial resolution in single-molecule fluorescence microscopy is discussed in light of its impact on catalytic assays. Key concepts are illustrated regarding the use of fluorescent reporters in catalytic reactions. Future challenges comprising the integration of other techniques, such as diffraction, scanning probe, or vibrational methods in single-molecule fluorescence spectroscopy are suggested.  相似文献   

6.
Many enzymes mold their structures to enclose substrates in their active sites such that conformational remodeling may be required during each catalytic cycle. In adenylate kinase (AK), this involves a large-amplitude rearrangement of the enzyme's lid domain. Using our method of high-resolution single-molecule FRET, we directly followed AK's domain movements on its catalytic time scale. To quantitatively measure the enzyme's entire conformational distribution, we have applied maximum entropy-based methods to remove photon-counting noise from single-molecule data. This analysis shows unambiguously that AK is capable of dynamically sampling two distinct states, which correlate well with those observed by x-ray crystallography. Unexpectedly, the equilibrium favors the closed, active-site-forming configurations even in the absence of substrates. Our experiments further showed that interaction with substrates, rather than locking the enzyme into a compact state, restricts the spatial extent of conformational fluctuations and shifts the enzyme's conformational equilibrium toward the closed form by increasing the closing rate of the lid. Integrating these microscopic dynamics into macroscopic kinetics allows us to model lid opening-coupled product release as the enzyme's rate-limiting step.  相似文献   

7.
Femtosecond time-resolved spectroscopy on model peptides with built-in light switches combined with computer simulation of light-triggered motions offers an attractive integrated approach toward the understanding of peptide conformational dynamics. It was applied to monitor the light-induced relaxation dynamics occurring on subnanosecond time scales in a peptide that was backbone-cyclized with an azobenzene derivative as optical switch and spectroscopic probe. The femtosecond spectra permit the clear distinguishing and characterization of the subpicosecond photoisomerization of the chromophore, the subsequent dissipation of vibrational energy, and the subnanosecond conformational relaxation of the peptide. The photochemical cis/trans-isomerization of the chromophore and the resulting peptide relaxations have been simulated with molecular dynamics calculations. The calculated reaction kinetics, as monitored by the energy content of the peptide, were found to match the spectroscopic data. Thus we verify that all-atom molecular dynamics simulations can quantitatively describe the subnanosecond conformational dynamics of peptides, strengthening confidence in corresponding predictions for longer time scales.  相似文献   

8.
We report fluorescence assays for a functionally important conformational change in bacteriophage T7 DNA polymerase (T7 pol) that use the environmental sensitivity of a Cy3 dye attached to a DNA substrate. An increase in fluorescence intensity of Cy3 is observed at the single-molecule level, reflecting a conformational change within the T7 pol ternary complex upon binding of a dNTP substrate. This fluorescence change is believed to reflect the closing of the T7 pol fingers domain, which is crucial for polymerase function. The rate of the conformational change induced by a complementary dNTP substrate was determined by both conventional stopped-flow and high-time-resolution continuous-flow fluorescence measurements at the ensemble-averaged level. The rate of this conformational change is much faster than that of DNA synthesis but is significantly reduced for noncomplementary dNTPs, as revealed by single-molecule measurements. The high level of selectivity of incoming dNTPs pertinent to this conformational change is a major contributor to replicative fidelity.  相似文献   

9.
A method for sensitively monitoring enzyme kinetics and activities by using dual-color fluorescence cross-correlation spectroscopy is described. This universal method enables the development of highly sensitive and precise assays for real-time kinetic analyses of any catalyzed cleavage or addition reaction, where a chemical linkage is formed or cleaved through an enzyme’s action between two fluorophores that can be discriminated spectrally. In this work, a homogeneous assay with restriction endonuclease EcoRI and a 66-bp double-stranded DNA containing the GAATTC recognition site and fluorophores at each 5′ end is described. The enzyme activity can be quantified down to the low picomolar range (>1.6 pM) where the rate constants are linearly dependent on the enzyme concentrations over two orders of magnitude. Furthermore, the reactions were monitored on-line at various initial substrate concentrations in the nanomolar range, and the reaction rates were clearly represented by the Michaelis–Menten equation with a KM of 14 ± 1 nM and a kcat of 4.6 ± 0.2 min−1. In addition to kinetic studies and activity determinations, it is proposed that enzyme assays based on the dual-color fluorescence cross-correlation spectroscopy will be very useful for high-throughput screening and evolutionary biotechnology.  相似文献   

10.
11.
The observation of subpicosecond fluctuations in the conformation of a small peptide in water is demonstrated. We use an experimental method that is specifically sensitive to conformational dynamics taking place on an ultrafast time scale. Complementary molecular-dynamics simulations confirm that the conformational fluctuations exhibit a subpicosecond component, the time scale and amplitude of which agree well with those derived from the experiment.  相似文献   

12.
Enzyme structural dynamics play a pivotal role in substrate binding and biological function, but the influence of substrate binding on enzyme dynamics has not been examined on fast time scales. In this work, picosecond dynamics of horseradish peroxidase (HRP) isoenzyme C in the free form and when ligated to a variety of small organic molecule substrates is studied by using 2D-IR vibrational echo spectroscopy. Carbon monoxide bound at the heme active site of HRP serves as a spectroscopic marker that is sensitive to the structural dynamics of the protein. In the free form, HRP assumes two distinct spectroscopic conformations that undergo fluctuations on a tens-of-picoseconds time scale. After substrate binding, HRP is locked into a single conformation that exhibits reduced amplitudes and slower time-scale structural dynamics. The decrease in carbon monoxide frequency fluctuations is attributed to reduced dynamic freedom of the distal histidine and the distal arginine, which are key residues in modulating substrate binding affinity. It is suggested that dynamic quenching caused by substrate binding can cause the protein to be locked into a conformation suitable for downstream steps in the enzymatic cycle of HRP.  相似文献   

13.
We have used chemical protein synthesis and advanced physical methods to probe dynamics-function correlations for the HIV-1 protease, an enzyme that has received considerable attention as a target for the treatment of AIDS. Chemical synthesis was used to prepare a series of unique analogues of the HIV-1 protease in which the flexibility of the "flap" structures (residues 37-61 in each monomer of the homodimeric protein molecule) was systematically varied. These analogue enzymes were further studied by X-ray crystallography, NMR relaxation, and pulse-EPR methods, in conjunction with molecular dynamics simulations. We show that conformational isomerization in the flaps is correlated with structural reorganization of residues in the active site, and that it is preorganization of the active site that is a rate-limiting factor in catalysis.  相似文献   

14.
Molecular chaperones are an essential part of the machinery that avoids protein aggregation and misfolding in vivo. However, understanding the molecular basis of how chaperones prevent such undesirable interactions requires the conformational changes within substrate proteins to be probed during chaperone action. Here we use single-molecule fluorescence spectroscopy to investigate how the DnaJ–DnaK chaperone system alters the conformational distribution of the denatured substrate protein rhodanese. We find that in a first step the ATP-independent binding of DnaJ to denatured rhodanese results in a compact denatured ensemble of the substrate protein. The following ATP-dependent binding of multiple DnaK molecules, however, leads to a surprisingly large expansion of denatured rhodanese. Molecular simulations indicate that hard-core repulsion between the multiple DnaK molecules provides the underlying mechanism for disrupting even strong interactions within the substrate protein and preparing it for processing by downstream chaperone systems.Maintaining protein homeostasis in vivo requires a tight regulation of protein folding to prevent misfolding and aggregation. Molecular chaperones have evolved as an essential part of the cellular machinery that facilitates such processes in the complex and crowded environment of a living cell (1, 2). To assist protein folding, many chaperones proceed through complex conformational cycles in an ATP-dependent manner (35). For several chaperone systems, these cycles have been investigated in great detail by experiment and simulation (68). A remarkable example are the heat shock protein (Hsp) 70 chaperones, which are essential in prokaryotes and eukaryotes and are involved in co-translational folding, refolding of misfolded and aggregated proteins, protein translocation, and protein degradation (9). The Hsp70 chaperone DnaK from Escherichia coli together with its co-chaperone DnaJ and the nucleotide exchange factor GrpE form an ATP-driven catalytic reaction cycle (7) (Fig. 1A). Many denatured or misfolded substrate proteins are first captured by DnaJ and subsequently transferred to the DnaK–ATP complex, with DnaK in an open conformation. Substrate and DnaJ synergistically trigger DnaK’s ATPase activity, which leads to locking of the substrate in the DnaK–ADP complex, with DnaK in the closed conformation. Driven by the following GrpE-catalyzed ADP–ATP exchange, the DnaK–substrate complex dissociates (10). Since this ATP-driven cycle can even solubilize protein aggregates (11, 12), substantial forces must be transduced to the substrate protein (1315). However, as for other chaperone systems (16), surprisingly little is known about how these forces and the resulting constraints of the underlying free energy surfaces affect the conformations of the denatured or misfolded substrate proteins. To better understand this important link between chaperone action and function, we probed the conformation of a substrate protein along the different stages of the chaperone cycle of DnaK with single-molecule Förster resonance energy transfer (smFRET), correlation spectroscopy, and microfluidic mixing.Open in a separate windowFig. 1.DnaK expands the denatured substrate protein. (A) Illustration of the DnaK–ATPase cycle. (B) Surface representation of rhodanese (PDB ID code 1RHS) with the subdomains indicated in different gray levels and the label positions of fluorescent dyes for single-molecule FRET measurements shown schematically. (C) FRET efficiency histograms of native rhodanese (gray) and denatured rhodanese under native conditions transiently populated in the microfluidic mixer (colored, measured 125 ms after dilution of rhodanese into native conditions). (D) FRET efficiency histograms of DnaJ–rhodanese complexes (0.5 µM DnaJ). (E) FRET efficiency histograms of DnaK–rhodanese complexes (0.5 µM DnaJ, 10 µM DnaK, and 1 mM ATP; DnaK and DnaJ were added simultaneously to rhodanese). Black lines indicate the DnaK–rhodanese complex population resulting from a fit that takes into account the residual population of refolded and DnaJ-bound rhodanese. The vertical lines in CE indicate the positions of the FRET efficiency peaks of the native population of the respective rhodanese variants. The small populations at zero transfer efficiency in D (note the axis scaling and the small amplitudes of this population compared with E) originate from incomplete elimination of molecules with inactive acceptor fluorophores by pulsed interleaved excitation.  相似文献   

15.
The recombination-activating gene products, RAG1 and RAG2, initiate V(D)J recombination during lymphocyte development by cleaving DNA adjacent to conserved recombination signal sequences (RSSs). The reaction involves DNA binding, synapsis, and cleavage at two RSSs located on the same DNA molecule and results in the assembly of antigen receptor genes. We have developed single-molecule assays to examine RSS binding by RAG1/2 and their cofactor high-mobility group-box protein 1 (HMGB1) as they proceed through the steps of this reaction. These assays allowed us to observe in real time the individual molecular events of RAG-mediated cleavage. As a result, we are able to measure the binding statistics (dwell times) and binding energies of the initial RAG binding events and characterize synapse formation at the single-molecule level, yielding insights into the distribution of dwell times in the paired complex and the propensity for cleavage on forming the synapse. Interestingly, we find that the synaptic complex has a mean lifetime of roughly 400 s and that its formation is readily reversible, with only ∼40% of observed synapses resulting in cleavage at consensus RSS binding sites.V(D)J recombination is responsible for assembling the variable regions of antigen receptor genes during B- and T-lymphocyte development. During V(D)J recombination, fragments of V, D, and J segments located together on particular chromosomes are rearranged into functional V(D)J or VJ alleles that are the specificity determinants for B-cell receptors or immunoglobulins (Igs) and T-cell receptors (TCR) (1). Adjacent to the V, D, and J gene segments are recombination signal sequences (RSSs) consisting of a conserved heptamer (consensus 5′-CACAGTG-3′) and nonamer (consensus 5′-ACAAAAACC-3′) separated by a spacer of 12 or 23 bp (referred to as 12RSS and 23RSS, respectively). Efficient recombination requires one 12RSS and one 23RSS, a constraint known as the 12/23 rule (1).The recombination-activating genes, RAG1 and RAG2, encode proteins that carry out V(D)J recombination by bringing a 12RSS and a 23RSS together into a paired (or synaptic) complex, nicking the RSS sites adjacent to the heptamer, and converting each nick into a double-strand break, leaving a hairpin at the coding end (Fig. 1). The hairpin is created by nucleophilic attack on the opposing strand by the 3′ hydroxyl group at the nick in a transesterification reaction. After RAG1/2 forms hairpins, it recruits the nonhomologous end joining machinery to repair the ends (1, 2). Since their discovery, the full-length RAG1 and RAG2 proteins have proven difficult to isolate and study in vitro (1). However, core domains (referred to as RAG1c and RAG2c) have been identified by removing a large region from the N terminus of RAG1 (which includes an E3 ubiquitin ligase domain) and a large region from the C terminus of RAG2 (which includes a plant homeodomain) (3, 4). These core proteins have been shown to tetramerize to form RAG1/2c, which retains RSS binding, nicking, and hairpin formation activities (5). High-mobility group-box protein 1 (HMGB1) acts as a cofactor and increases RAG1/2c affinity for the 23RSS (6). HMGB1 is required for paired complex formation and efficient conversion of RAG-mediated nicks into hairpins (710).Open in a separate windowFig. 1.Orchestration of V(D)J cleavage. (A) Schematic of key molecules in the V(D)J recombination process and nomenclature used throughout the paper. (B) During the cleavage phase of V(D)J recombination, RAG1/2c and HMGB1 bind and nick a 12RSS (magenta) and a 23RSS (yellow) spaced by a distance known as the intersignal distance. Then, the two RSSs are brought together, forming a looped paired complex. While in the paired complex state, hairpins are produced only in the presence of a 12RSS and a 23RSS, which is known as the 12/23 rule. *The stoichiometry of RAG1/2c and HMGB1 involved in forming the paired complex is unknown.Current in vitro assays that capture the paired complex with RAG1/2c generally place a 12RSS and a 23RSS on two different DNA molecules, typically short oligonucleotides. However, in vivo, antigen receptor loci are assembled using RSSs on the same DNA molecule (1, 11). One prior study captured RAG1/2c and HMGB1 bound to DNA with a 12RSS and a 23RSS on the same substrate using standard bulk assays (12). However, many key mechanistic questions remain unresolved concerning how a diverse immune repertoire is generated by the RAG recombinase. How long does an RAG complex spend bound to a 12RSS or a 23RSS? What is the lifetime of an RAG-mediated paired complex formed between a 12RSS and a 23RSS before DNA double-strand break (hairpin) formation? Is formation of this complex reversible, or does it inevitably go on to cleavage? These issues have implications for the mechanisms that determine Ig and TCR repertoires.To address these questions, we have explored the dynamics of various stages of the V(D)J recombination cleavage reaction on single DNA molecules in real time, allowing visualization and characterization of individual RAG–HMGB1–RSS complexes and paired complex formation using intersignal distances comparable with some of the shorter distances found in the assembly of V, D, and J gene segments in vivo (13). Using a tethered particle motion (TPM) assay (14), we observed RSS-dependent apparent shortening of DNA in the presence of RAG1/2c, probably a reflection of bending of the DNA, and were able to exploit this shortening as a signature of the RAG1/2c–RSS-bound state (Fig. 2A). This technique in conjunction with a statistical mechanical model of RAG binding have allowed us to determine single-molecule binding constants for RAG1/2c with or without HMGB1 (Fig. 2 A and B). These measurements are also used to determine the dwell time distributions for the 12RSS and 23RSS protein bound states. Furthermore, this same substrate length readout is used for direct detection of the paired complex before DNA cleavage (Fig. 2C), because bringing the two RSSs into close proximity effectively shortens the tether. We find that formation of the paired complex results in one of two outcomes: either the DNA is cleaved, causing bead release from its tether, or the complex dissociates, leaving the DNA template intact. Finally, we use bead loss on cleavage to quantify the time dependence of cleavage on the concentrations of RAG1/2c and HMGB1 (Fig. 2D). These experiments provide a quantitative picture of the stages leading up to DNA cleavage during V(D)J recombination and reveal the dynamic nature of the paired complex.Open in a separate windowFig. 2.Steps of V(D)J cleavage investigated by TPM in this study: RAG binding, RAG–HMGB1 binding, paired complex formation, and hairpin production. Schematic of constructs and experimental measurements to examine the V(D)J recombination cleavage reaction at various stages. (A) RAG1/2c binding to the 12RSS or the 23RSS site alone. (B) RAG1/2c binding to the 12RSS or the 23RSS site in the presence of HMGB1. (C) RAG1/2c binding to DNA substrates containing 12RSS and 23RSS for the purposes of observing formation of a looped paired complex. (D) Bead loss caused by DNA cleavage as a result of 12/23 rule-regulated hairpin production.  相似文献   

16.
Molecular chaperones are known to be essential for avoiding protein aggregation in vivo, but it is still unclear how they affect protein folding mechanisms. We use single-molecule Förster resonance energy transfer to follow the folding of a protein inside the GroEL/GroES chaperonin cavity over a time range from milliseconds to hours. Our results show that confinement in the chaperonin decelerates the folding of the C-terminal domain in the substrate protein rhodanese, but leaves the folding rate of the N-terminal domain unaffected. Microfluidic mixing experiments indicate that strong interactions of the substrate with the cavity walls impede the folding process, but the folding hierarchy is preserved. Our results imply that no universal chaperonin mechanism exists. Rather, a competition between intra- and intermolecular interactions determines the folding rates and mechanisms of a substrate inside the GroEL/GroES cage.  相似文献   

17.
The eukaryotic signaling protein calmodulin (CaM) can bind to more than 300 known target proteins to regulate numerous functions in our body in a calcium-dependent manner. How CaM distinguishes between these various targets is still largely unknown. Here, we investigate fluctuations of the complex formation of CaM and its target peptide sequences using single-molecule force spectroscopy by AFM. By applying mechanical force, we can steer a single CaM molecule through its folding energy landscape from the fully unfolded state to the native target-bound state revealing equilibrium fluctuations between numerous intermediate states. We find that the prototypical CaM target sequence skMLCK, a fragment from skeletal muscle myosin light chain kinase, binds to CaM in a highly cooperative way, while only a lower degree of interdomain binding cooperativity emerges for CaMKK, a target peptide from CaM-dependent kinase kinase. We identify minimal binding motifs for both of these peptides, confirming that affinities of target peptides are not exclusively determined by their pattern of hydrophobic anchor residues. Our results reveal an association mode for CaMKK in which the peptide binds strongly to only partially Ca2+-saturated CaM. This binding mode might allow for a fine-tuning of the intracellular response to changes in Ca2+ concentration.  相似文献   

18.
A method was developed to monitor dynamic changes in protein structure and interfacial behavior on surfaces by single-molecule Förster resonance energy transfer. This method entails the incorporation of unnatural amino acids to site-specifically label proteins with single-molecule Förster resonance energy transfer probes for high-throughput dynamic fluorescence tracking microscopy on surfaces. Structural changes in the enzyme organophosphorus hydrolase (OPH) were monitored upon adsorption to fused silica (FS) surfaces in the presence of BSA on a molecule-by-molecule basis. Analysis of >30,000 individual trajectories enabled the observation of heterogeneities in the kinetics of surface-induced OPH unfolding with unprecedented resolution. In particular, two distinct pathways were observed: a majority population (∼ 85%) unfolded with a characteristic time scale of 0.10 s, and the remainder unfolded more slowly with a time scale of 0.7 s. Importantly, even after unfolding, OPH readily desorbed from FS surfaces, challenging the common notion that surface-induced unfolding leads to irreversible protein binding. This suggests that protein fouling of surfaces is a highly dynamic process because of subtle differences in the adsorption/desorption rates of folded and unfolded species. Moreover, such observations imply that surfaces may act as a source of unfolded (i.e., aggregation-prone) protein back into solution. Continuing study of other proteins and surfaces will examine whether these conclusions are general or specific to OPH in contact with FS. Ultimately, this method, which is widely applicable to virtually any protein, provides the framework to develop surfaces and surface modifications with improved biocompatibility.Understanding the effect of near-surface environments on protein conformation is critical in many bioengineering and biomedical applications, including biosensing, cell culture, tissue engineering, biocatalysis, and pharmaceutical formulation. Importantly, surface interactions that perturb protein structure can inactivate proteins, as has widely been observed in the case of surface-immobilized enzymes (16). Such interactions, by inducing unfolding and subsequent accumulation of freely absorbing proteins on biomaterial surfaces, may trigger unfavorable cellular responses (710). However, experimental methods to elucidate both protein structure and interfacial dynamics (e.g., adsorption, diffusion, desorption), particularly in heterogeneous near-surface environments, are virtually nonexistent.Conventional methods to determine surface effects on protein structure are largely ensemble-averaging techniques that provide limited mechanistic insight. Such limitations can lead to misinterpretations of apparent interfacial phenomena, which are used to explain the biocompatibility of biomaterials. For example, conventional biophysical methods often find that the average surface protein conformation relaxes from native-like to non-native-like upon surface adsorption, which is interpreted as evidence for unfolding (1114). However, the time scales over which these conformational changes reportedly occur are orders of magnitude longer than typical surface residence times of isolated proteins (15), suggesting that the picture of surface-induced spreading may be oversimplified. In particular, this interpretation overlooks alternative hypotheses, including that the surface may “collect” nonnative protein molecules or that the surface may indirectly influence protein–protein interactions, causing aggregation. In part, ambiguity stems from the fact that molecules can continuously absorb and desorb, complicating the analysis of surface-induced unfolding kinetics. In addition, such methods are unable to capture heterogeneous behavior, which can arise from statistical variation in protein structure, surface aggregation, or irregularities in the underlying surface.Recent advances in dynamic single-molecule (SM) fluorescence microscopy techniques present considerable opportunities for structure determination of adsorbed and immobilized proteins. Importantly, by measuring the entire distribution of protein structure (as opposed to only the ensemble average), such techniques can better capture the effects of protein heterogeneity. For example, using SM total internal reflection fluorescence microscopy (TIRFM), Kastantin and colleagues (15) uncovered the dynamic behavior of fibrinogen on biomaterial surfaces, showing that fibrinogen–surface interactions were highly transient and, moreover, that fibrinogen oligomers diffused more slowly on surfaces and had longer residence times than monomers. Furthermore, when intermolecular interactions were maximized by formation of an organized protein layer, residence times increased (16). These works suggested that protein aggregation, which may be mediated by surface chemistry, plays a prominent role in the formation of stable protein layers.This article describes a unique approach to characterizing dynamic changes in protein structure in near-surface environments using dynamic SM microscopy. Specifically, this approach exploits SM-Förster resonance energy transfer (FRET) tracking to elucidate changes in protein structure at the SM level. To enable SM resolution, this approach was used in conjunction with protein engineering, using unnatural amino acids as a means to site-specifically introduce donor and acceptor fluorophores. Importantly, by incorporating donor and acceptor fluorophores at different sites via protein engineering, interfacial effects on local protein structure may be precisely probed. Structural changes in different regions of a protein also may be combined with dynamic interfacial, as well as functional, measurements, allowing direct conclusions to be drawn with respect to ways in which protein conformation, dynamics, and activity are connected. Although site-specific labeling methods have been used with FRET to monitor conformational changes in single protein molecules in bulk solution (1719) or covalently tethered to surfaces (2022), such methods have not been applied to the analysis of proteins undergoing dynamic adsorption, desorption, and diffusion while interacting directly with material interfaces. SM-FRET tracking, with temporal resolution of 100 ms, has previously been used to monitor dynamic changes in conformation of mobile DNA at the solution–surface interface (23, 24), using high-throughput tracking algorithms that permit the observation of 104–106 molecules.The utility of SM-FRET tracking microscopy, when combined with protein engineering, for surface characterization was demonstrated by studying the interaction of the enzyme organophosphorus hydrolase (OPH) with FS. OPH, which catalyzes the hydrolysis of organophosphate compounds, is structurally and functionally well-characterized and is also of interest because of the relevance of OPH–surface interactions in the development of biosensors and self-decontaminating surfaces for sensing and destroying toxic nerve agents and pesticides (25, 26). The unnatural amino acid p-azido-l-phenylalanine (AzF) was incorporated in OPH so that donor and acceptor fluorophores could be site-specifically attached. Analysis of FRET as a function of time after adsorption enabled characterization of surface-induced unfolding kinetics of OPH and the fate of the unfolded OPH. SM-FRET tracking revealed that the mean unfolding time of individual molecules is 0.19 s and that 99.99% of molecules resided on the surface for less than 30 s. In contrast, ensemble unfolding analysis found that formation of an unfolded layer of OPH on FS required longer than 9 h. This profound discrepancy suggests that the mechanisms behind apparent surface-induced protein unfolding and layer formation should be reconsidered.  相似文献   

19.
A theory for single-molecule fluorescence detection is developed and then used to analyze data from subpicomolar solutions of B-phycoerythrin (PE). The distribution of detected counts is the convolution of a Poissonian continuous background with bursts arising from the passage of individual fluorophores through the focused laser beam. The autocorrelation function reveals single-molecule events and provides a criterion for optimizing experimental parameters. The transit time of fluorescent molecules through the 120-fl imaged volume was 800 microseconds. The optimal laser power (32 mW at 514.5 nm) gave an incident intensity of 1.8 x 10(23) photons.cm-2.s-1, corresponding to a mean time of 1.1 ns between absorptions. The mean incremental count rate was 1.5 per 100 microseconds for PE monomers and 3.0 for PE dimers above a background count rate of 1.0. The distribution of counts and the autocorrelation function for 200 fM monomer and 100 fM dimer demonstrate that single-molecule detection was achieved. At this concentration, the mean occupancy was 0.014 monomer molecules in the probed volume. A hard-wired version of this detection system was used to measure the concentration of PE down to 1 fM. This single-molecule counter is 3 orders of magnitude more sensitive than conventional fluorescence detection systems.  相似文献   

20.
The equilibrium folding of the catalytic domain of Bacillus subtilis RNase P RNA is investigated by single-molecule fluorescence resonance energy transfer (FRET). Previous ensemble studies of this 255-nucleotide ribozyme described the equilibrium folding with two transitions, U-to-I(eq)-to-N, and focused on the I(eq)-to-N transition. The present study focuses on the U-to-I(eq) transition. Comparative ensemble measurements of the ribozyme construct labeled with fluorescein at the 5' end and Cy3 at the 3' end show that modifications required for labeling do not interfere with folding and help to define the Mg(2+) concentration range for the U-to-I(eq) transition. Histogram analysis of the Mg(2+)-dependent single-molecule FRET efficiency reveals two previously undetermined folding intermediates. The single-molecule FRET trajectories exhibit non-two-state and nonergodic behaviors at intermediate Mg(2+) concentrations on the time scale of seconds. The trajectories at intermediate Mg(2+) concentrations are classified into five classes based on three FRET levels and their dynamics of interconversion within the measured time range. This heterogeneity, together with the observation of "nonsudden jump" FRET transitions, indicates that the early folding steps of this ribozyme involve a series of intermediates with different degrees of kinetic isolation and that folding occurs under kinetic control and involves many "local" conformational switches. A free energy contour is constructed to illustrate the complex folding surface.  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