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 62 毫秒
1.
Prion and prion-like diseases involve the propagation of misfolded protein conformers. Small-molecule pharmacological chaperones can inhibit propagated misfolding, but how they interact with disease-related proteins to prevent misfolding is often unclear. We investigated how pentosan polysulfate (PPS), a polyanion with antiprion activity in vitro and in vivo, interacts with mammalian prion protein (PrP) to alter its folding. Calorimetry showed that PPS binds two sites on natively folded PrP, but one PPS molecule can bind multiple PrP molecules. Force spectroscopy measurements of single PrP molecules showed PPS stabilizes not only the native fold of PrP but also many different partially folded intermediates that are not observed in the absence of PPS. PPS also bound tightly to unfolded segments of PrP, delaying refolding. These observations imply that PPS can act through multiple possible modes, inhibiting misfolding not only by stabilizing the native fold or sequestering natively folded PrP into aggregates, as proposed previously, but also by binding to partially or fully unfolded states that play key roles in mediating misfolding. These results underline the likely importance of unfolded states as critical intermediates on the prion conversion pathway.

The propagation of misfolded proteins is a hallmark of a wide variety of diseases, including Alzheimer’s disease, Parkinson’s disease, amyotrophic lateral sclerosis, and the tauopathies (1). In misfolding diseases like these, nonnative conformers of disease-linked proteins can convert natively folded molecules into additional copies of the misfolded conformers, propagating the misfolding (2). The archetypal misfolding diseases, where propagated misfolding was first discovered, are the prionopathies, such as Creutzfeldt-Jakob disease (CJD) in humans, scrapie in sheep, “mad-cow” disease, and chronic wasting disease in cervids. In prion diseases, the helical fold of native cellular prion protein, PrPC, is converted to a β-rich misfolded isoform, PrPSc, that is both toxic to neurons and acts as an infectious agent to propagate the disease (3).Because misfolding and conversion are central to prion and prion-like diseases, one therapeutic strategy has been to search for small molecules that can act as pharmacological “chaperones” to inhibit the misfolding (4), analogous to the chaperone proteins that help suppress misfolding in the cell (5). Even though the structure of PrPSc and the mechanism by which it converts PrPC remain controversial (3, 6), a variety of chemically distinct compounds with promising potential against prions have been found (7, 8). The mechanisms by which these putative chaperones act are typically quite poorly understood, however, in part because it is challenging to observe the effects of antiprion compounds on PrP folding directly. As a result, it has proven difficult to develop effective therapeutics or indeed to use the effects of antiprion compounds to gain insight into the misfolding mechanism.Single-molecule methods are well suited to address this challenge because they can observe individual molecules directly as they change conformation and distinguish subpopulations in a heterogeneous mixture (9, 10), and they have been applied to study misfolding in several disease-related proteins (1116), including PrP (1719). Such approaches have also been used to study the mechanisms of cellular chaperones (20), but the effects of pharmacological chaperones have been little explored at the single-molecule level. The only study to date (21) examined the changes in the folding of PrP caused by binding of an antiprion iron tetrapyrrole, Fe-TMPyP, finding that it not only stabilized the native fold as proposed based on ensemble studies (22) but, more interestingly, that it also inhibited the interactions that drove the formation of stable misfolded aggregates, similar to the action of cellular chaperones (23). Owing to the paucity of single-molecule studies of pharmacological chaperones, most of the existing proposals for their mechanism of action are based on the limited information available from ensemble assays, and it remains unclear how they achieve their effect.Here, we examine how the antiprion compound pentosan polysulfate (PPS) changes the folding of PrP. PPS belongs to a different class of pharmacological chaperone from Fe-TMPyP and other similar molecules: Instead of having a compact, planar structure with cationic charge and low molecular weight, PPS has the opposite character—it is a large, linear polymer with a heparin-like structure (Fig. 1 A, Inset) and anionic charge. It has been found to inhibit the accumulation of PrPSc in prion-infected cells and prolong the lifespan of infected animals (24), and it has even had some success increasing the survival period of patients with CJD, although it is difficult to administer, as it does not penetrate the blood–brain barrier (2528). PPS is known to interact with PrP (7, 29) and has been proposed to sequester PrPC in nontoxic aggregates so that it cannot be converted to PrPSc (7, 30), but its mechanism of action remains uncertain. To study the interaction between PPS and PrP in more detail, we used a combination of isothermal titration calorimetry (ITC), measuring the thermodynamics of the interaction, and single-molecule force spectroscopy (SMFS), observing the dynamics of a single PrP molecule as it was unfolded and refolded while held under tension by optical tweezers in the presence of PPS. We found very heterogeneous binding: not only binding to two independent sites on natively folded PrP, where each PPS molecule interacted with multiple PrP molecules, but also strong binding to partially and fully unfolded states. The fact that two chemically distinct antiprion agents both interact strongly with unfolded states suggests a central role for unfolded PrP in the prion conversion mechanism.Open in a separate windowFig. 1.ITC of PPS binding to PrPC. (A) ITC thermogram for binding of PPS to hamster PrPC at pH 7. (Inset) Structure of PPS. (B) Binding isotherm (black) and fit to model with two independent binding sites (red).  相似文献   

2.
Chronic wasting disease (CWD) is a relentless epidemic disorder caused by infectious prions that threatens the survival of cervid populations and raises increasing public health concerns in North America. In Europe, CWD was detected for the first time in wild Norwegian reindeer (Rangifer tarandus) and moose (Alces alces) in 2016. In this study, we aimed at comparing the strain properties of CWD prions derived from different cervid species in Norway and North America. Using a classical strain typing approach involving transmission and adaptation to bank voles (Myodes glareolus), we found that prions causing CWD in Norway induced incubation times, neuropathology, regional deposition of misfolded prion protein aggregates in the brain, and size of their protease-resistant core, different from those that characterize North American CWD. These findings show that CWD prion strains affecting Norwegian cervids are distinct from those found in North America, implying that the highly contagious North American CWD prions are not the proximate cause of the newly discovered Norwegian CWD cases. In addition, Norwegian CWD isolates showed an unexpected strain variability, with reindeer and moose being caused by different CWD strains. Our findings shed light on the origin of emergent European CWD, have significant implications for understanding the nature and the ecology of CWD in Europe, and highlight the need to assess the zoonotic potential of the new CWD strains detected in Europe.

Prion diseases, otherwise known as transmissible spongiform encephalopathies (TSEs), are invariably fatal infectious neurodegenerative disorders of animals and humans. Prions are composed largely or exclusively of self-replicating protein aggregates of PrPSc, which is a misfolded isoform of the cellular prion protein (PrPC). Most human prion diseases, including Creutzfeldt–Jakob disease (CJD) and sporadic fatal insomnia, occur sporadically and are thought to result from the spontaneous misfolding of PrPC into PrPSc. Mutations in the coding sequence of the gene encoding PrP (PRNP) in inherited diseases such as familial CJD, fatal familial insomnia, and Gerstmann–Sträussler–Scheinker syndrome are thought to potentiate spontaneous misfolding of PrPC to PrPSc (1). The intrinsic transmissibility of prions has also resulted in epidemics of acquired prion diseases (2), with significant public health consequences. Examples include iatrogenic CJD transmission caused by prion-contaminated human growth hormone preparations and dura mater grafts, and variant CJD resulting from human exposure to bovine spongiform encephalopathy (BSE) prions.Animal prion diseases also occur as contagious forms giving rise to outbreaks such as classical scrapie in small ruminants and chronic wasting disease (CWD) in cervids (3). In addition to their impact on human health, exemplified by zoonotic BSE transmission (2), epidemic animal prion diseases, such as CWD in North America (NA) (4), have significant ecologic and economic consequences. First identified in captive deer in Northern Colorado in the 1960s, CWD is now known to affect wild and farmed cervid populations in 26 states and three Canadian provinces. Disease was introduced to the Republic of Korea following importation of subclinical elk (Cervus canadensis) from Canada (5). The irrevocable spread of CWD in NA raises serious concerns about the survival of cervid populations, as well as for the risks posed to other animals and to humans (6).Prion diseases are caused by different prion strains, which are thought to be encoded by conformational variants of PrPSc (7, 8). While different PrPSc conformers can be discriminated using biochemical approaches, assessing infectious properties of prions in susceptible hosts remains the gold standard for strain characterization. Such analyses are of the utmost importance for trace-back studies of iatrogenic (9, 10) or zoonotic prion strains (11, 12), for identifying spillover hosts (13), for estimating the zoonotic potential of classical and atypical scrapie strains (14, 15), and for understanding the epidemiology and evolution of strains in contagious prion diseases (1620).In Europe, CWD was detected for the first time in wild Norwegian reindeer (Rangifer tarandus) and moose (Alces alces) in 2016 (21, 22). The emergence of CWD in Europe necessitates characterization of the strain properties of these CWD prions of uncertain origin to explore a possible link with NA CWD. Knowledge about Norwegian CWD strains is essential to assess their risk to animal and human health, and to develop evidence-based policies to control and limit the spread of the disease (23).In a previous study, we found that bank voles carrying isoleucine at codon 109 (Bv109I) are highly susceptible to CWD isolates from the United States. After subpassage in Bv109I, the same vole-adapted CWD strain was isolated from NA mule deer (Odocoileus hemionus), white-tailed deer (Odocoileus virginianus), and elk isolates, which was characterized by an unprecedented short disease duration and peculiar and recognizable strain features (24). Here, we extended our previous characterizations of NA CWD isolates and compared the biological properties of CWD isolates from Norway and NA by transmission to the Bv109I genetic line of bank voles.  相似文献   

3.
Peripheral myelin protein (PMP22) is an integral membrane protein that traffics inefficiently even in wild-type (WT) form, with only 20% of the WT protein reaching its final plasma membrane destination in myelinating Schwann cells. Misfolding of PMP22 has been identified as a key factor in multiple peripheral neuropathies, including Charcot-Marie-Tooth disease and Dejerine–Sottas syndrome. While biophysical analyses of disease-associated PMP22 mutants show altered protein stabilities, leading to reduced surface trafficking and loss of PMP22 function, it remains unclear how destabilization of PMP22 mutations causes mistrafficking. Here, native ion mobility–mass spectrometry (IM-MS) is used to compare the gas phase stabilities and abundances for an array of mutant PM22 complexes. We find key differences in the PMP22 mutant stabilities and propensities to form homodimeric complexes. Of particular note, we observe that severely destabilized forms of PMP22 exhibit a higher propensity to dimerize than WT PMP22. Furthermore, we employ lipid raft–mimicking SCOR bicelles to study PMP22 mutants, and find that the differences in dimer abundances are amplified in this medium when compared to micelle-based data, with disease mutants exhibiting up to 4 times more dimer than WT when liberated from SCOR bicelles. We combine our findings with previous cellular data to propose that the formation of PMP22 dimers from destabilized monomers is a key element of PMP22 mistrafficking.

The misfolding of membrane proteins is implicated in the mechanisms of multiple debilitating diseases such as cystic fibrosis and retinitis pigmentosa (14). Specific membrane protein mutations are often associated with disease states, with variant forms exhibiting altered stability and cellular trafficking (5). Unfortunately, due to the challenges associated with preparing and handling pure, highly concentrated membrane protein samples, detailed structural information on such targets is often lacking, especially for disease mutant forms. Furthermore, as some membrane proteins associated with misfolding-based diseases have hundreds of mutations of interest (3), there is a clear need for high-throughput methods to assess disease mutation-induced changes in membrane protein stability and structure.Peripheral myelin protein 22 (PMP22) is such a membrane protein, for which misfolding and trafficking of mutant variants have been implicated in disease (6). PMP22 is a tetra-span integral membrane glycoprotein predominately expressed in Schwann cells, which are the principal glial cells of the peripheral nervous system (PNS), where they produce myelin (79). In addition to accounting for ∼5% of the protein found in the myelin sheath surrounding PNS nerve axons, PMP22 is thought to regulate intracellular Ca2+ levels (10), apoptosis (11), linkage of the actin cytoskeleton with lipid rafts (12), formation of epithelial intercellular junctions (13), myelin formation (14), lipid metabolism, and cholesterol trafficking (15). Dysregulation and misfolding of PMP22 has been identified as a key factor in multiple neurodegenerative disorders, such as Charcot-Marie-Tooth disease types 1A and E, as well as Dejerine–Sottas syndrome (6, 1618). Like a number of other disease-linked membrane proteins (19), the trafficking of PMP22 is known to be inefficient, with only 20% of the wild-type (WT) protein reaching its final plasma membrane destination in Schwann cells (16, 20). Previously, it has been shown through a range of biophysical analyses that disease-associated PMP22 mutations lower thermodynamic protein stability as the root cause of reduced trafficking and loss of protein function; however, the mechanism by which destabilization of PMP22 causes mistrafficking is still not well understood (6). Additionally, a high-resolution structure of PMP22 has not yet been published.Native mass spectrometry (MS) has recently been demonstrated to overcome sample purity and concentration barriers to reveal critical details of membrane protein structure and function (2123). Through the use of nano-electrospray (nESI), intact membrane proteins are ionized within detergent micelles or other membrane mimetics (2427), which can then be removed from the membrane protein ions within the instrument. This method has been used to elucidate oligomeric state (2830), complex organization (31, 32), and lipid interactions (3335) of diverse membrane proteins. The addition of ion-mobility separation–mass spectrometry (IM-MS) provides data on the orientationally averaged size of analytes (36) and enables collision induced unfolding (CIU) experiments (37). In CIU, the energies experienced by gas-phase protein ions are increased in a stepwise fashion causing gas-phase protein unfolding to occur. These dynamic measurements have been shown to be sensitive to ligand binding (38, 39), glycosylation (40, 41), and disulfide bonding (40) in soluble proteins, as well as selective lipid and small molecule binding in membrane proteins (4245). While CIU can clearly capture subtle structural changes in membrane proteins (43, 45, 46) and soluble mutant protein variants (47, 48) its ability to characterize membrane protein variants is only beginning to be explored.Here, we demonstrate the ability of native MS and CIU to detect key differences in the gas-phase stability and homodimer complex formation of PMP22 variants, together leading to insights into the mechanism of PMP22 dysregulation in disease. We quantify the propensity of PMP22 to dimerize across WT and seven disease-associated point mutations. We find that mutations associated with severe disease states form significantly more dimer than WT. Through CIU, we quantify the stability of gas-phase monomeric and dimeric PMP22 and find that variants bearing mutations associated with severe neuropathy exhibit the lowest relative monomer conformational stability. Interestingly, we also observe that dimers formed by various disease mutant forms of PMP22 are all more stable than WT PMP22 dimeric complexes. We continue by comparing our results to previously published biophysical datasets and find that our monomeric PMP22 gas-phase stability values correlate well with cellular trafficking data (6). Finally, we probe the effects of solubilization agents on PMP22 by characterizing its dimerization within sphingomyelin and cholesterol rich (SCOR) bicelles (49). We find that dimeric PMP22 complexes persist within SCOR bicelles and that the mutants resulting in the most severe disease phenotypes form higher population of dimer than WT. We conclude by describing a possible mechanism of PMP22 dysregulation in severe neurodegenerative diseases by which PMP22 monomers are destabilized, leading to dimers that traffic much less efficiently to the plasma membrane than WT PMP22.  相似文献   

4.
5.
Robust estimates for the rates and trends in terrestrial gross primary production (GPP; plant CO2 uptake) are needed. Carbonyl sulfide (COS) is the major long-lived sulfur-bearing gas in the atmosphere and a promising proxy for GPP. Large uncertainties in estimating the relative magnitude of the COS sources and sinks limit this approach. Sulfur isotope measurements (34S/32S; δ34S) have been suggested as a useful tool to constrain COS sources. Yet such measurements are currently scarce for the atmosphere and absent for the marine source and the plant sink, which are two main fluxes. Here we present sulfur isotopes measurements of marine and atmospheric COS, and of plant-uptake fractionation experiments. These measurements resulted in a complete data-based tropospheric COS isotopic mass balance, which allows improved partition of the sources. We found an isotopic (δ34S ± SE) value of 13.9 ± 0.1‰ for the troposphere, with an isotopic seasonal cycle driven by plant uptake. This seasonality agrees with a fractionation of −1.9 ± 0.3‰ which we measured in plant-chamber experiments. Air samples with strong anthropogenic influence indicated an anthropogenic COS isotopic value of 8 ± 1‰. Samples of seawater-equilibrated-air indicate that the marine COS source has an isotopic value of 14.7 ± 1‰. Using our data-based mass balance, we constrained the relative contribution of the two main tropospheric COS sources resulting in 40 ± 17% for the anthropogenic source and 60 ± 20% for the oceanic source. This constraint is important for a better understanding of the global COS budget and its improved use for GPP determination.

The Earth system is going through rapid changes as the climate warms and CO2 level rises. This rise in CO2 is mitigated by plant uptake; hence, it is important to estimate global and regional photosynthesis rates and trends (1). Yet, robust tools for investigating these processes at a large scale are scarce (2). Recent studies suggest that carbonyl sulfide (COS) could provide an improved constraint on terrestrial photosynthesis (gross primary production, GPP) (212). COS is the major long-lived sulfur-bearing gas in the atmosphere and the main supplier of sulfur to the stratospheric sulfate aerosol layer (13), which exerts a cooling effect on the Earth’s surface and regulates stratospheric ozone chemistry (14).During terrestrial photosynthesis, COS diffuses into leaf stomata and is consumed by photosynthetic enzymes in a similar manner to CO2 (35). Contrary to CO2, COS undergoes rapid and irreversible hydrolysis mainly by the enzyme carbonic-anhydrase (6, 7). Thus, COS can be used as a proxy for the one-way flux of CO2 removal from the atmosphere by terrestrial photosynthesis (2, 811). However, the large uncertainties in estimating the COS sources weaken this approach (1012, 15). Tropospheric COS has two main sources: the oceans and anthropogenic emissions, and one main sink–terrestrial plant uptake (8, 1013). Smaller sources include biomass burning, soil emissions, wetlands, volcanoes, and smaller sinks include OH destruction, stratospheric destruction, and soil uptake (12). The largest source of COS to the atmosphere is the ocean, both as direct COS emission, and as indirect carbon disulfide (CS2) and dimethylsulfide (DMS) emissions that are rapidly oxidized to COS (10, 1620). Recent studies suggest oceanic COS emissions are in the range of 200–4,000 GgS/y (1922). The second major COS source is the anthropogenic source, which is dominated by indirect emissions derived from CS2 oxidation, mainly from the use of CS2 as an industrial solvent. Direct emissions of COS are mainly derived from coal and fuel combustion (17, 23, 24). Recent studies suggest that anthropogenic emissions are in the range of 150–585 GgS/y (23, 24). The terrestrial plant uptake is estimated to be in the range of 400–1,360 GgS/y (11). Measurements of sulfur isotope ratios (δ34S) in COS may be used to track COS sources and thus reduce the uncertainties in their flux estimations (15, 2527). However, the isotopic mass balance approach works best if the COS end members are directly measured and have a significantly different isotopic signature. Previous δ34S measurements of atmospheric COS are scarce and there have been no direct measurements of two important components: the δ34S of oceanic COS emissions, and the isotopic fractionation of COS during plant uptake (15, 2527). In contrast to previous studies that used assessments for these isotopic values, our aim was to directly measure the isotopic values of these missing components, and to determine the tropospheric COS δ34S variability over space and time.  相似文献   

6.
Cell–cell adhesions are often subjected to mechanical strains of different rates and magnitudes in normal tissue function. However, the rate-dependent mechanical behavior of individual cell–cell adhesions has not been fully characterized due to the lack of proper experimental techniques and therefore remains elusive. This is particularly true under large strain conditions, which may potentially lead to cell–cell adhesion dissociation and ultimately tissue fracture. In this study, we designed and fabricated a single-cell adhesion micro tensile tester (SCAµTT) using two-photon polymerization and performed displacement-controlled tensile tests of individual pairs of adherent epithelial cells with a mature cell–cell adhesion. Straining the cytoskeleton–cell adhesion complex system reveals a passive shear-thinning viscoelastic behavior and a rate-dependent active stress-relaxation mechanism mediated by cytoskeleton growth. Under low strain rates, stress relaxation mediated by the cytoskeleton can effectively relax junctional stress buildup and prevent adhesion bond rupture. Cadherin bond dissociation also exhibits rate-dependent strengthening, in which increased strain rate results in elevated stress levels at which cadherin bonds fail. This bond dissociation becomes a synchronized catastrophic event that leads to junction fracture at high strain rates. Even at high strain rates, a single cell–cell junction displays a remarkable tensile strength to sustain a strain as much as 200% before complete junction rupture. Collectively, the platform and the biophysical understandings in this study are expected to build a foundation for the mechanistic investigation of the adaptive viscoelasticity of the cell–cell junction.

Adhesive organelles between neighboring epithelial cells form an integrated network as the foundation of complex tissues (1). As part of normal physiology, this integrated network is constantly exposed to mechanical stress and strain, which is essential to normal cellular activities, such as proliferation (24), migration (5, 6), differentiation (7), and gene regulation (7, 8) associated with a diverse set of functions in tissue morphogenesis (911) and wound healing (9). A host of developmental defects or clinical pathologies in the form of compromised cell–cell associations will arise when cells fail to withstand external mechanical stress due to genetic mutations or pathological perturbations (12, 13). Indeed, since the mechanical stresses are mainly sustained by the intercellular junctions, which may represent the weakest link and limit the stress tolerance within the cytoskeleton network of a cell sheet, mutations or disease-induced changes in junction molecules and components in adherens junctions and desmosomes lead to cell layer fracture and tissue fragility, which exacerbate the pathological conditions (1417). This clinical relevance gives rise to the importance of understanding biophysical transformations of the cell–cell adhesion interface when cells are subjected to mechanical loads.As part of their normal functions, cells often experience strains of tens to a few hundred percent at strain rates of 10−4 to 1 s−1 (1821). For instance, embryonic epithelia are subjected to strain rates in the range of 10−4 to 10−3 s−1 during normal embryogenesis (22). Strain rates higher than 0.1 s−1 are often experienced by adult epithelia during various normal physiological functions (21, 23, 24), such as breathing motions in the lung (1 to 10 s−1) (25), cardiac pulses in the heart (1 to 6.5 s−1) (20), peristaltic movements in the gut (0.4 to 1.5 s−1), and normal stretching of the skin (0.1 to 5 s−1). Cells have different mechanisms to dissipate the internal stress produced by external strain to avoid fracture, often via cytoskeleton remodeling and cell–cell adhesion enhancement (26, 27). These coping mechanisms may have different characteristic timescales. Cytoskeleton remodeling can dissipate mechanical stress promptly due to its viscoelastic nature and the actomyosin-mediated cell contractility (17, 2832). Adhesion enhancement at the cell–cell contact is more complex in terms of timescale. Load-induced cell–cell adhesion strengthening has been shown via the increase in the number of adhesion complexes (3335) or by the clustering of adhesion complexes (3639), which occurs on a timescale ranging from a few minutes up to a few hours after cells experience an initial load (28). External load on the cell–cell contact also results in a prolonged cell–cell adhesion dissociation time (40, 41), suggesting cadherin bonds may transition to catch bonds under certain loading conditions (42, 43), which can occur within seconds (44). With the increase in cellular tension, failure to dissipate the stress within the cell layer at a rate faster than the accumulation rate will inevitably lead to the fracture of the cell layer (45). Indeed, epithelial fracture often aggravates the pathological outcomes in several diseases, such as acute lung injuries (46), skin disorders (47), and development defects (48). It is generally accepted that stress accumulation in the cytoskeleton network (49, 50) and potentially in the cytoplasm is strain-rate–dependent (51). However, to date, there is a lack of understanding about the rate-dependent behavior of cell–cell adhesions, particularly about which of the stress-relaxation mechanisms are at play across the spectrum of strain rates. In addition, it remains unclear how the stress relaxation interplays with adhesion enhancement under large strains, especially at high strain rates which may lead to fracture, that is, a complete separation of mature cell–cell adhesions under a tensile load (45, 52, 53). Yet, currently, there is a lack of quantitative technology that enables the investigation of these mechanobiological processes in a precisely controlled manner. This is especially true at high strain rates.To delineate this mechanical behavior, the cleanest characterization method is to directly measure stress dynamics at a single mature cell–cell adhesion interface. Specifically, just as a monolayer cell sheet is a reduction from three-dimensional (3D) tissue, a single cell–cell adhesion interface, as a reduction from a monolayer system, represents the smallest unit to study the rheological behavior of cellular junctions. The mechanistic understanding uncovered with this single unit will inform cellular adaptations to a more complex stress microenvironment in vivo and in vitro, in healthy and diseased conditions. To this end, we developed a single-cell adhesion micro tensile tester (SCAµTT) platform based on nanofabricated polymeric structures using two-photon polymerization (TPP). This platform allows in situ investigation of stress–strain characteristics of a mature cell–cell junction through defined strains and strain rates. With SCAµTT, we reveal some interesting biophysical phenomena at the single cell–cell junction that were previously not possible to observe using existing techniques. We show that cytoskeleton growth can effectively relax intercellular stress between an adherent cell pair in a strain-rate–dependent manner. Along with cadherin-clustering–induced bond strengthening, it prevents failure to occur at low strain rates. At high strain rates, insufficient relaxation leads to stress accumulation, which results in cell–cell junction rupture. We show that a remarkably large strain can be sustained before junction rupture (>200%), even at a strain rate as high as 0.5 s−1. Collectively, the rate-dependent mechanical characterization of the cell–cell junction builds the foundation for an improved mechanistic understanding of junction adaptation to an external load and potentially the spatiotemporal coordination of participating molecules at the cell–cell junction.  相似文献   

7.
One of the hallmarks of Alzheimer’s disease and several other neurodegenerative disorders is the aggregation of tau protein into fibrillar structures. Building on recent reports that tau readily undergoes liquid–liquid phase separation (LLPS), here we explored the relationship between disease-related mutations, LLPS, and tau fibrillation. Our data demonstrate that, in contrast to previous suggestions, pathogenic mutations within the pseudorepeat region do not affect tau441’s propensity to form liquid droplets. LLPS does, however, greatly accelerate formation of fibrillar aggregates, and this effect is especially dramatic for tau441 variants with disease-related mutations. Most important, this study also reveals a previously unrecognized mechanism by which LLPS can regulate the rate of fibrillation in mixtures containing tau isoforms with different aggregation propensities. This regulation results from unique properties of proteins under LLPS conditions, where total concentration of all tau variants in the condensed phase is constant. Therefore, the presence of increasing proportions of the slowly aggregating tau isoform gradually lowers the concentration of the isoform with high aggregation propensity, reducing the rate of its fibrillation. This regulatory mechanism may be of direct relevance to phenotypic variability of tauopathies, as the ratios of fast and slowly aggregating tau isoforms in brain varies substantially in different diseases.

Tau is a major neuronal protein that plays a key role in Alzheimer’s disease (AD) and a number of other neurodegenerative disorders that are collectively classified as tauopathies. The latter include frontotemporal dementia with parkinsonism linked to chromosome 17 (FTDP-17), progressive supranuclear palsy, Pick’s disease, corticobasal degeneration, and chronic traumatic encephalopathy (15). Under normal physiological conditions, tau is localized to axons where it is involved in the assembly of microtubules (16). In tauopathies, the protein self-associates into different forms of filaments that contain largely hyperphosphorylated tau and have properties of amyloid fibrils (15).Alternative splicing of the MAPT gene that encodes tau results in six major isoforms in the human central nervous system. These isoforms differ with respect to the number of N-terminal inserts as well as the number of 31 to 32 residue pseudorepeat sequences in the C-terminal part of the protein (15). Structurally, tau is largely an intrinsically disordered protein, with local secondary structures existing only within the pseudorepeat region (1, 7). A large number of mutations have been identified in the latter region that correlate with inherited cases of FTDP-17 (8, 9). These mutations not only diminish the ability of tau to promote microtubule assembly, but many also promote self-association of tau into amyloid fibrils (1012). This strongly suggests that tau misfolding and aggregation is one of the key events in disease pathogenesis.A number of recent reports indicate that purified full-length tau (tau441) has a high propensity to undergo liquid–liquid phase separation (LLPS) in vitro in the presence of crowding agents that emulate the high concentration of macromolecules in the cell. This was observed both for the phosphorylated (13) and nonphosphorylated protein (1416), and it was determined that tau LLPS is driven largely by attractive electrostatic intermolecular interactions between the negatively charged N-terminal and positively charged middle/C-terminal regions of the protein (15). Tau condensation into droplets (complex coacervation) was also observed in the presence of polyanions such as RNA or heparin (17, 18). These observations in vitro are partially supported by studies in cells (13, 1924), especially within the context of tau interaction with microtubules (21). However, it remains unclear whether tau could undergo LLPS in cells on its own or, rather, its recruitment to membraneless organelles such as stress granules is largely driven by interactions with other proteins and/or RNA. These limitations notwithstanding, the observations that tau has a propensity for LLPS have potentially important implications for the pathogenic process in tauopathies, as studies with other proteins involved in neurodegenerative diseases (e.g., TDP-43, FUS) indicate that the environment of liquid droplets is conducive to the pathological aggregation of these proteins (2532). In line with these findings, it was recently suggested that LLPS can initiate tau aggregation. However, the evidence for this was very limited and largely based on optical microscopy observations (13).In the present study, we explored the relationship between pathogenic mutations of tau, protein LLPS, and aggregation into amyloid fibrils. Our data show that, in contrast to previous suggestions (13), pathogenic mutations within the pseudorepeat region do not affect the propensity of tau to undergo LLPS. These mutations, however, do dramatically accelerate the liquid-to-solid phase transition within the droplets, leading to rapid formation of fibrillar aggregates. Most important, this study also reveals a previously unrecognized mechanism by which LLPS can regulate the rate of amyloid formation in mixtures containing tau isoforms with different aggregation propensities. These findings strongly suggest that LLPS may play a major regulatory role in the formation of pathological tau aggregates in neurodegenerative diseases.  相似文献   

8.
Electrochemical CO2 or CO reduction to high-value C2+ liquid fuels is desirable, but its practical application is challenged by impurities from cogenerated liquid products and solutes in liquid electrolytes, which necessitates cost- and energy-intensive downstream separation processes. By coupling rational designs in a Cu catalyst and porous solid electrolyte (PSE) reactor, here we demonstrate a direct and continuous generation of pure acetic acid solutions via electrochemical CO reduction. With optimized edge-to-surface ratio, the Cu nanocube catalyst presents an unprecedented acetate performance in neutral pH with other liquid products greatly suppressed, delivering a maximal acetate Faradaic efficiency of 43%, partial current of 200 mA⋅cm−2, ultrahigh relative purity of up to 98 wt%, and excellent stability of over 150 h continuous operation. Density functional theory simulations reveal the role of stepped sites along the cube edge in promoting the acetate pathway. Additionally, a PSE layer, other than a conventional liquid electrolyte, was designed to separate cathode and anode for efficient ion conductions, while not introducing any impurity ions into generated liquid fuels. Pure acetic acid solutions, with concentrations up to 2 wt% (0.33 M), can be continuously produced by employing the acetate-selective Cu catalyst in our PSE reactor.

Electrochemically reducing carbon dioxide (CO2) or carbon monoxide (CO) to commodity fuels or chemicals, with the input of green and economical renewable electricity, has become an alternative route to traditional chemical engineering processes (17). To date, a variety of catalysts have been demonstrated to reduce CO2 into C1 products (CO, formate, etc.) with high Faradaic efficiencies (FEs) (6, 814); however, the critical step of C-C coupling toward high-value C2+ products was rarely observed on catalysts other than Cu (1520). Due to its proper binding strength with *CO (19, 21, 22), Cu has shown the capability of catalyzing the coupling of two carbons or more in CO2/CO reduction reactions (CO2RR/CORR), which, however, typically suffers from large overpotential and low selectivity (2325). In previous studies on Cu catalysts, tremendous efforts have been focused on how to promote C-C coupling toward high selectivity of C2+ products, including facet engineering (2628), grain boundary engineering (29, 30), surface modifications (3133), etc. For example, recent studies on facet dependence of C-C coupling have shown that the C2+ selectivity could be improved on Cu(100) compared to other facets due to more stabilized *OCCO intermediates (26, 34). While the overall C2+ selectivity is constantly improved by researchers in this field, the generated products are typically a mixture of different C2+ compounds, which subsequently necessitates energy- and cost-intensive downstream product separation and purification processes (3537). A solution to continuously obtain a highly pure, single C2+ product stream, which can be directly fed into practical applications, will impose a significant scientific and economic impact but has been rarely studied before.Ethylene, ethanol, and acetate are the three major C2 products that have been widely reported in CO2RR/CORR on Cu materials (1618, 3844). Compared to gas-phase products, liquid products show significant advantages due to their high energy densities and ease of storage and distribution (45). Nevertheless, the generation of a high-purity C2 liquid fuel via electrochemical CO2RR/CORR is challenging due to the involvement of two types of impurities. First, in previous studies of Cu catalysts, while the C2 product can be dramatically increased by strategies such as exposing (100) facets, ethanol and acetic acid usually share the same potential window and are cogenerated simultaneously, making it difficult to obtain a single C2 liquid product (29, 46, 47); Second, in traditional H-cell or flow-cell reactors, the generated liquid products were mixed with solutes in liquid electrolytes such as KOH or KHCO3, which requires extra separation and concentration processes to recover pure liquid fuel solutions in practical applications (37, 48, 49). An integration of rational design in both catalytic material and reactor, for high single-C2 selectivity and electrolyte-free liquid fuel output, respectively, therefore becomes the key to achieve the goal of pure C2 liquid fuels.Here we report the continuous generation of pure acetic acid solutions via CORR on Cu nanocube (NC) catalysts in a porous solid electrolyte (PSE) reactor. By flexibly tuning the edge to (100) surface ratio, the medium-size Cu NC catalyst exhibits a maximal acetate Faradaic efficiency of 43% with a partial current density of ∼200 mA⋅cm−2 in neutral pH, setting up a different acetate performance benchmark (18, 40, 41). More importantly, the selectivity of other liquid products (ethanol and trace amount of n-propanol) was suppressed to below 2%, suggesting an ultrahigh acetic acid relative purity of up to 98 wt%. An impressive stability was demonstrated by a continuous operation of CORR under 150 mA⋅cm−2 current for over 150 h with negligible degradation in selectivity and activity. This acetate-selective Cu NC catalyst was successfully employed into our PSE reactor for the continuous generation of electrolyte-free, pure acetic acid solutions, with overall current of up to 1 A⋅cm−2 and acetic acid relative purity as high as 96 wt%. Different from the traditional liquid electrolyte, our PSE layer can efficiently conduct ions while it does not introduce any impurity ions into the generated liquid products. Density functional theory (DFT) calculations suggest that the stepped edge sites on the Cu NC readily dissociate *OCCOH into *CCO and *OH, which is a key step in promoting the formation of acetate. We propose that the large NC has a low edge-to-facet ratio and therefore cogenerates considerable ethanol and ethylene instead of acetate, while a small NC provides less active (100) area for C-C coupling and therefore predominantly catalyzes the hydrogen evolution reaction (HER) instead of yielding C2 products.  相似文献   

9.
L-type voltage-gated CaV1.2 channels crucially regulate cardiac muscle contraction. Activation of β-adrenergic receptors (β-AR) augments contraction via protein kinase A (PKA)–induced increase of calcium influx through CaV1.2 channels. To date, the full β-AR cascade has never been heterologously reconstituted. A recent study identified Rad, a CaV1.2 inhibitory protein, as essential for PKA regulation of CaV1.2. We corroborated this finding and reconstituted the complete pathway with agonist activation of β1-AR or β2-AR in Xenopus oocytes. We found, and distinguished between, two distinct pathways of PKA modulation of CaV1.2: Rad dependent (∼80% of total) and Rad independent. The reconstituted system reproduces the known features of β-AR regulation in cardiomyocytes and reveals several aspects: the differential regulation of posttranslationally modified CaV1.2 variants and the distinct features of β1-AR versus β2-AR activity. This system allows for the addressing of central unresolved issues in the β-AR–CaV1.2 cascade and will facilitate the development of therapies for catecholamine-induced cardiac pathologies.

Cardiac excitation–contraction coupling crucially depends on the L-type voltage-dependent Ca2+ channel, CaV1.2. Influx of extracellular Ca2+ via CaV1.2 triggers Ca2+ release from the sarcoplasmic reticulum via the Ca2+ release channel (1). Activation of the sympathetic nervous system increases heart rate, relaxation rate and contraction force. The latter is largely due to increased Ca2+ influx via CaV1.2 (2, 3). Pathological prolonged sympathetic activation progressively impairs cardiac function, causing heart failure, partly due to misregulation of CaV1.2 (4, 5).Cardiac CaV1.2 is a heterotrimer comprising the pore-forming subunit α1C (∼240 kDa), the intracellular CaVβ2 (∼68 kDa) and the extracellular α2δ (∼170 kDa) (Fig. 1A) (6, 7). The N and C termini (NT, CT respectively) of α1C are cytosolic and vary among CaV1.2 isoforms. Further, most of the cardiac α1C protein is posttranslationally cleaved at the CT, around amino acid (a.a.) 1800, to produce the truncated ∼210-kDa α1C protein and the ∼35-kDa cleaved distal CT (dCT); however, the full-length protein is also present (811).Open in a separate windowFig. 1.cAMP regulation of CaV1.2 is enhanced by coexpression of Rad. (A) CaV1.2 and Rad. α1C and α2δ subunits are shown schematically, with structures of β2b (38) and Rad (74). The truncation in α1CΔ1821 was at a.a. 1,821 (red cross mark) similar to naturally truncated cardiac α1C, ∼a.a. 1800 (9). CaVβ binds to the cytosolic loop I, L1, that connects repeat domains I and II. Rad exerts inhibitory action on the channel, in part through an interaction with CaVβ. (B) Rad reduces the Ba2+ current of CaV1.2-α1CΔ1821 (α1CΔ1821, β2b and α2δ; 1.5 ng RNA of each subunit) in a dose-dependent manner. Pearson correlation, r = −0.82, P = 0.023. Each point represents mean ± SEM from 7 to 10 oocytes recorded during 1 d. The linear regression line was drawn for nonzero doses of Rad. (C) Rad enhances the cAMP-induced increase in IBa. Diary plots of the time course of change in IBa (normalized to initial IBa) are shown before and after intracellular injection of cAMP in representative cells. No Rad: Upper; with Rad: Lower. (Insets) Currents at +20 mV before (black trace) and 10 min after cAMP injection (red trace). (D) “before–after” plots of cAMP-induced changes in IBa in individual cells injected Rad RNA while varying Rad:β2b RNA ratio (by weight, wt/wt). Empty symbols–before cAMP; red-filled–after cAMP. n = 3 experiments; statistics: paired t test. (E) cAMP-induced increase in IBa at different Rad/β2b RNA levels (summary of data from D). Each symbol represents fold increase in IBa induced by cAMP injection in one cell. Here and in the following figures, box plots show 25 to 75 percentiles, whiskers show the 5/95 percentiles, and black and red horizontal lines within the boxes are the median and mean, respectively. At all Rad:β2b RNA ratios except 1:20, the cAMP-induced increase in IBa was significantly greater than without Rad (Kruskal–Wallis test; H = 36.1, 6 degrees of freedom, P < 0.001). (F) Summary of cAMP effects in 10 experiments without and with Rad at 1:2 and 1:1 Rad:β2b RNA ratios (pooled). Number of cells: within the bars. Statistics: Mann–Whitney U test; U = 19.0, P < 0.001.The sympathetic nervous system activates cardiac β-adrenergic receptors (β-AR), primarily β1-AR (which is coupled to Gs, is globally distributed in cardiomyocytes, and mediates most of the β-AR-enhancement of contraction and CaV1.2 activity) and β2-AR, which can couple to both Gs and Gi (12). The cascade of adrenergic modulation of CaV1.2 comprises agonist binding to β-ARs, activation of Gs and adenylyl cyclase, elevated intracellular cAMP levels, and activation of protein kinase A (PKA) by cAMP-induced dissociation of its catalytic subunit (PKA-CS) from the regulatory subunit. However, the final step, how PKA-CS enhances CaV1.2 activity, remained enigmatic. A long-standing paradigm was a direct phosphorylation by PKA-CS of α1C and/or CaVβ subunits (3, 1316). However, numerous studies critically challenged this theory. In particular, mutated CaV1.2 channels in genetically engineered mice lacking putative PKA phosphorylation sites on α1C and/or β2b, were still up-regulated by PKA (9, 1721) (reviewed in refs. 6 and 22).One significant obstacle in deciphering the mechanism of PKA regulation of CaV1.2 was a recurrent lack of success in reconstituting the regulation in heterologous systems, which proved challenging and controversial (23). Studies in heterologous cellular models, including Xenopus oocytes, demonstrated that cAMP failed to up-regulate CaV1.2 containing the full-length α1C, CaV1.2-α1C (2426). However, robust β-AR–induced up-regulation of Ca2+ currents was observed in oocytes injected with total heart RNA (27, 28), suggesting the necessity of an auxiliary protein, the “missing link” (24, 25). Interestingly, partial regulation was observed with dCT-truncated α1C (16, 29). Intracellular injection of cAMP or PKA-CS in Xenopus oocytes caused a modest (30 to 40%) up-regulation of CaV1.2, containing a dCT-truncated α1C, CaV1.2-α1CΔ1821 (29). This regulation required the presence of the initial segment of the long-NT of α1C but did not involve CaVβ subunit. We proposed that this mechanism might account for part of the adrenergic regulation of CaV1.2 in the heart (29). Normally adrenergic stimulation in cardiomyocytes increases the Ca2+ current two- to threefold; thus, a major part of the regulation has remained unexplained.Recently, Liu et al. identified Rad as the “missing link” in PKA regulation of CaV1.2 (20). Rad is a member of the Ras-related GTP-binding protein subfamily (RGK) that inhibit high voltage-gated calcium channels CaV1 and CaV2 (30). Rad tonically inhibits CaV1.2, largely via an interaction with CaVβ (31, 32). Ablation of Rad in murine heart was shown to increase basal CaV1.2 activity and rendered the channel insensitive to β-AR regulation, probably through a “ceiling” effect (33, 34). Liu et al. (20) reconstituted a major part of the CaV1.2 regulation cascade, initiated by forskolin-activated adenylyl cyclase in mammalian cells, ultimately attaining an approximately twofold increase in Ca2+ current. The regulation required phosphorylation of Rad, the presence of CaVβ, and the interaction of CaVβ with the cytosolic loop I of α1C, suggesting that PKA phosphorylation of Rad reduces its interaction with CaVβ and relieves the tonic inhibition of CaV1.2 (20, 35).Importantly, the complete adrenergic cascade, starting with β-AR activation, has not yet been heterologously reconstituted for CaV1.2. Also, the relation between the Rad-dependent regulation and the regulation reported in our previous study (29) is not clear. Here, we utilized the Xenopus oocyte heterologous expression system and successfully reconstituted the entire β-AR cascade. We demonstrate two distinct pathways of PKA modulation of CaV1.2 (Rad dependent and Rad independent) and characterize the roles of NT and CT of α1C, β2b, and Rad in the adrenergic modulation of cardiac CaV1.2 channels. Reproducing the complete β-AR cascade in a heterologous expression system will promote the identification and characterization of intracellular proteins that regulate the cascade, eventually assisting efforts to develop therapies to treat heart failure and other catecholamine-induced cardiac pathologies.  相似文献   

10.
Learning and memory are assumed to be supported by mechanisms that involve cholinergic transmission and hippocampal theta. Using G protein–coupled receptor-activation–based acetylcholine sensor (GRABACh3.0) with a fiber-photometric fluorescence readout in mice, we found that cholinergic signaling in the hippocampus increased in parallel with theta/gamma power during walking and REM sleep, while ACh3.0 signal reached a minimum during hippocampal sharp-wave ripples (SPW-R). Unexpectedly, memory performance was impaired in a hippocampus-dependent spontaneous alternation task by selective optogenetic stimulation of medial septal cholinergic neurons when the stimulation was applied in the delay area but not in the central (choice) arm of the maze. Parallel with the decreased performance, optogenetic stimulation decreased the incidence of SPW-Rs. These findings suggest that septo–hippocampal interactions play a task-phase–dependent dual role in the maintenance of memory performance, including not only theta mechanisms but also SPW-Rs.

The neurotransmitter acetylcholine is thought to be critical for hippocampus-dependent declarative memories (1, 2). Reduction in cholinergic neurotransmission, either in Alzheimer’s disease or in experiments with cholinergic antagonists, such as scopolamine, impairs memory function (38). Acetylcholine may bring about its beneficial effects on memory encoding by enhancing theta rhythm oscillations, decreasing recurrent excitation, and increasing synaptic plasticity (911). Conversely, drugs which activate cholinergic receptors enhance learning and, therefore, are a neuropharmacological target for the treatment of memory deficits in Alzheimer’s disease (5, 12, 13).The contribution of cholinergic mechanisms in the acquisition of long-term memories and the role of the hippocampal–entorhinal–cortical interactions are well supported by experimental data (5, 12, 13). In addition, working memory or “short-term” memory is also supported by the hippocampal–entorhinal–prefrontal cortex (1416). Working memory in humans is postulated to be a conscious process to “keep things in mind” transiently (16). In rodents, matching to sample task, spontaneous alternation between reward locations, and the radial maze task have been suggested to function as a homolog of working memory [“working memory like” (17)].Cholinergic activity is a critical requirement for working memory (18, 19) and for sustaining theta oscillations (10, 2022). In support of this contention, theta–gamma coupling and gamma power are significantly higher in the choice arm of the maze, compared with those in the side arms where working memory is no longer needed for correct performance (2326). It has long been hypothesized that working memory is maintained by persistent firing of neurons, which keep the presented items in a transient store in the prefrontal cortex and hippocampal–entorhinal system (2731), although the exact mechanisms are debated (3237). An alternative hypothesis holds that items of working memory are stored in theta-nested gamma cycles (38). Common in these models of working memory is the need for an active, cholinergic system–dependent mechanism (3941). However, in spontaneous alternation tasks, the animals are not moving continuously during the delay, and theta oscillations are not sustained either. During the immobility epochs, theta is replaced by intermittent sharp-wave ripples (SPW-R), yet memory performance does not deteriorate. On the contrary, artificial blockade of SPW-Rs can impair memory performance (42, 43), and prolongation of SPW-Rs improves performance (44). Under the cholinergic hypothesis of working memory, such a result is unexpected.To address the relationship between cholinergic/theta versus SPW-R mechanism in spontaneous alternation, we used a G protein–coupled receptor-activation–based acetylcholine sensor (GRABACh3.0) (45) to monitor acetylcholine (ACh) activity during memory performance in mice. In addition, we optogenetically enhanced cholinergic tone, which suppresses SPW-Rs by a different mechanism than electrically or optogenetically induced silencing of neurons in the hippocampus (43, 44). We show that cholinergic signaling in the hippocampus increases in parallel with theta power/score during walking and rapid eye movement (REM) sleep and reaches a transient minimum during SPW-Rs. Selective optogenetic stimulation of medial septal cholinergic neurons decreased the incidence of SPW-Rs during non-REM sleep (4648), as well as during the delay epoch of a working memory task and impaired memory performance. These findings demonstrate that memory performance is supported by complementary theta and SPW-R mechanisms.  相似文献   

11.
12.
Ras is regulated by a specific guanine nucleotide exchange factor Son of Sevenless (SOS), which facilitates the exchange of inactive, GDP-bound Ras with GTP. The catalytic activity of SOS is also allosterically modulated by an active Ras (Ras–GTP). However, it remains poorly understood how oncogenic Ras mutants interact with SOS and modulate its activity. Here, native ion mobility–mass spectrometry is employed to monitor the assembly of the catalytic domain of SOS (SOScat) with KRas and three cancer-associated mutants (G12C, G13D, and Q61H), leading to the discovery of different molecular assemblies and distinct conformers of SOScat engaging KRas. We also find KRasG13D exhibits high affinity for SOScat and is a potent allosteric modulator of its activity. A structure of the KRasG13D•SOScat complex was determined using cryogenic electron microscopy providing insight into the enhanced affinity of the mutant protein. In addition, we find that KRasG13D–GTP can allosterically increase the nucleotide exchange rate of KRas at the active site more than twofold compared to KRas–GTP. Furthermore, small-molecule Ras•SOS disruptors fail to dissociate KRasG13D•SOScat complexes, underscoring the need for more potent disruptors. Taken together, a better understanding of the interaction between oncogenic Ras mutants and SOS will provide avenues for improved therapeutic interventions.

Ras, a member of the small G-protein family, represents important signaling molecules with diverse cellular roles, such as cell differentiation and proliferation (14). Different isoforms of Ras (HRas, KRas, and NRas) have high overall sequence identity and are the most commonly mutated of all discovered oncogenes (5, 6). Of the three Ras isoforms, KRas is the most frequently mutated isoform in cancers, such as pancreatic cancer (70–90%), colon cancer (30–50%), and lung cancer (20–30%) (7, 8). Ras proteins regulate cell signaling pathways by cycling between inactive, GDP-bound, and active, GTP-bound states that is accompanied by remodeling of three key regions within Ras: p-loop (residues 10–17), switch I (residues 30–38), and switch II (residues 60–76) (911).As Ras proteins possess slow guanine nucleotide exchange rates, their activation is regulated by guanine nucleotide exchange factors (GEFs) that reload Ras with GTP (12, 13). The multidomain protein, Son of Sevenless (SOS), is a GEF with the cdc25 and Ras exchanger motif domains representing the minimal, functionally competent unit, termed SOScat (14). Structural studies have revealed two Ras binding sites to SOScat, leading to the discovery that binding of Ras–GTP at the distal (or allosteric) site allosterically modulates SOScat activity, which markedly increases the nucleotide exchange rate at the active site (14, 15). In addition, the degree of this allosteric modulation greatly depends on the nucleotide-bound state of Ras (14, 16, 17). Moreover, SOS is conformationally dynamic and binding of Ras at the allosteric site appears to shift the population to active conformation(s) of SOS (18). In addition, SOScat samples a broad range of turnover rates by fluctuating between distinct, long-lived functional states (19). Despite these advances, the nucleotide specificity of Ras bound to the active and allosteric sites of SOS and assembly with RAS, including oncogenic mutants, at the molecular level is poorly understood.Targeting oncogenic Ras mutants presents significant challenges because of their relatively smooth surface that lacks potentially druggable pockets (6). Nevertheless, the discovery of KRas inhibitors, particularly those that form irreversible covalent bonds with Cys-12, comprises one of the most active areas of cancer research (6, 20). As there are few windows of opportunity to specifically target Ras mutants (6), apart from covalent binding to Cys-12, other approaches have also been explored, such as designing molecules to disrupt Ras•SOS interactions, thereby preventing activation of Ras (20). An increasing number of small-molecule disruptors and peptide mimetics have been designed to disrupt the Ras•SOS interaction (21). Potent small-molecule disruptors have recently been discovered that inhibit the formation of the Ras•SOS complex and demonstrate antiproliferative activity, representing a viable approach for targeting Ras-driven tumors (22, 23). These results highlight the therapeutic importance of disrupting the interaction between Ras and SOS.Over the past three decades, native ion mobility–mass spectrometry (IM-MS) has evolved as a powerful analytical technique to investigate protein complexes and their interaction with other molecules (2426). In native MS, biological samples in aqueous ammonium acetate are ionized using nanoelectrospray ionization and introduced into a mass spectrometer tuned to preserve noncovalent interactions and structure (24, 27). Native IM-MS can provide information on protein complexes, such as subunit stoichiometry and topology (28), and, unlike other biophysical techniques, resolve individual ligand-binding events (29). In combination with an apparatus to control temperature, native MS can be used to determine equilibrium binding constants and thermodynamics for protein–ligand and protein–protein that are in direct agreement with traditional biophysical approaches, such as isothermal calorimetry (3035). Recently, native MS has been employed to determine transition state thermodynamics for the intrinsic GTPase activity of KRas and several oncogenic mutants (36). Notably, intrinsic GTPase activity rates determined using traditional solution-based assays mirrored those obtained using native MS (36).Although the interaction of Ras and SOS has been the subject of numerous studies, the interaction of oncogenic Ras mutants with SOS remains poorly described. To better understand the role of these interactions in cancer, native IM-MS is used to characterize the molecular assemblies formed between SOScat and mutants of KRas associated with cancer. IM spectrometry shows conformational heterogeneity of SOScat with specific conformers engaging KRas. Three selected oncogenic mutants of KRas form distinct molecular assemblies with SOScat, such as KRasG13D forming exclusively a ternary complex with SOScat. The cryogenic electron microscopy (cryo-EM) structure of the KRasG13D•SOScat complex provides insight into the mechanism for the higher affinity of KRasG13D for SOScat. KRasG13D–GTP also allosterically modulates the activity of SOScat more than the wild-type protein. In addition, the recent inhibitors developed to disrupt the KRas•SOScat complex, BAY-293 and BI-3406, cannot dissociate KRasG13D•SOScat complexes. Other small molecules, such as ARS-1620 and Kobee0065, display a range of efficacies in disrupting complexes formed between KRas mutants and SOScat.  相似文献   

13.
The timescale for the microscopic dynamics of proteins during conformational transitions is set by the intrachain diffusion coefficient, D. Despite the central role of protein misfolding and aggregation in many diseases, it has proven challenging to measure D for these processes because of their heterogeneity. We used single-molecule force spectroscopy to overcome these challenges and determine D for misfolding of the prion protein PrP. Observing directly the misfolding of individual dimers into minimal aggregates, we reconstructed the energy landscape governing nonnative structure formation. Remarkably, rather than displaying multiple pathways, as typically expected for aggregation, PrP dimers were funneled into a thermodynamically stable misfolded state along a single pathway containing several intermediates, one of which blocked native folding. Using Kramers’ rate theory, D was found to be 1,000-fold slower for misfolding than for native folding, reflecting local roughening of the misfolding landscape, likely due to increased internal friction. The slow diffusion also led to much longer transit times for barrier crossing, allowing transition paths to be observed directly for the first time to our knowledge. These results open a new window onto the microscopic mechanisms governing protein misfolding.The formation of intricate 3D structures by proteins is a complex physical process. Such “folding” is typically described in terms of energy landscape theory (1) as a thermally driven diffusive search over an energy landscape in conformational space for the minimum-energy structure. In this picture, whereas the rates at which structural transitions take place are dominated by the presence of energy barriers in the landscape (2), it is the coefficient of diffusion over the landscape, D, that encapsulates the microscopic dynamics of the protein chain, setting the characteristic timescale for molecular motions. Knowledge of D provides insight into the internal friction in the protein chain as it undergoes conformational fluctuations (3) and sets the ultimate speed limit at which changes in structure can take place (4).Given the fundamental importance of the diffusion coefficient in protein folding, there has been much interest in measuring D under different conditions. Conformational diffusion has been studied extensively in peptides and unfolded proteins (510), using fluorescence probes such as fluorophore quenching or Förster resonant energy transfer to measure reconfiguration times. Typically, D ∼107–108 nm2/s was found, although values as low as 105 nm2/s have been reported (10). Because the diffusion coefficient is inversely proportional to friction, measurements of D have been important for investigating the role and origin of internal friction along the folding pathway (6, 9). Possible links between the value of D and aggregation propensity have also been explored in intrinsically disordered proteins (5). However, it has proven challenging to measure the diffusion coefficient during barrier crossing via fluorescence, owing to the very brief transition time for barrier crossing (11).Recently, an alternate approach using single-molecule force spectroscopy (SMFS), whereby force is applied to induce structural changes in an isolated molecule, has been applied to measure D for barrier crossings (1215). This approach takes advantage of the powerful ability of force spectroscopy to measure energy landscapes by analyzing the statistics of conformational fluctuations (16), using the landscape profile and rates to recover D from Kramers’ theory (17). The ability to probe barrier crossings opens up the exciting possibility of investigating the microscopic differences between native structure formation and what happens when folding goes awry, producing nonnative structures. Such misfolding is a feature of many diseases, with misfolded proteins characteristically aggregating into insoluble amyloid fibers rich in β-sheets (18). The value of D should be particularly important in misfolding, because the outcome of misfolding and aggregation processes is thought in many cases to be dominated by kinetics (19). Although in silico studies have begun to probe the energy landscapes for protein misfolding and aggregation (20), these landscapes have not yet been reconstructed experimentally. A direct comparison between diffusion in native folding and misfolding of the same protein has therefore not yet been made.An important challenge in studying misfolding and aggregation is that they are typically very heterogeneous processes, involving various transient species along multiple pathways leading to different types of aggregates (21). Despite recent advances including solving the structures of some native-like aggregation precursor states (22, 23) and small oligomers (24), a full picture of the sequence of molecular events in aggregation remains elusive. Single-molecule assays are well suited to overcome this challenge, through their ability to distinguish and characterize even transient subpopulations with high sensitivity (25). Previous studies have investigated phenomena ranging from transient, metastable misfolding events (2628) to various stages in the growth of aggregates (29, 30), but complete misfolding pathways leading to stable misfolded states have not yet been elucidated (25).Here we use SMFS to reconstruct the energy landscape for misfolding of the prion protein PrP and thereby recover D for misfolding. PrP is particularly interesting as a model for investigating protein misfolding, because even though PrP folds rapidly into its native structure (PrPC) under normal conditions (31), it has a stable misfolded form, PrPSc, which can convert PrPC to form more PrPSc, thereby acting as an infectious agent to transmit prion diseases (32). The structure of PrPSc remains unknown, as does the mechanism for conversion of PrPC, although a variety of structural models (33) and mechanisms (32, 34) have been proposed. Previous single-molecule studies of PrP misfolding (35) have characterized properties such as the kinetics of oligomerization (30) and the dependence of aggregation pathways on metal ions (36), suggesting that the conversion of PrP features several phases starting with dimerization, but detailed pathways were not resolved. SMFS measurements of isolated PrP monomers found that they frequently sampled various misfolded conformations, but they were not thermodynamically stable, consistent with the view that misfolded PrP is stable only within aggregates (27).Because the conversion of PrP may start with dimerization (30, 37), here we have focused on dimers as the smallest form of aggregate. Using optical tweezers to apply tension and thereby unfold and refold single PrP dimers, we found that dimers readily misfold into stable aggregates. A single misfolding pathway was observed and characterized in detail, revealing several exceptional features that set PrP apart from other proteins studied to date. Reconstruction of the energy landscape for dimer misfolding, along with the previously measured landscape for native folding (15), was then used to make the first comparison of D for native folding versus misfolding in the same protein, showing that diffusion is much slower during misfolding than during native folding.  相似文献   

14.
15.
Spatiotemporal regulation of signaling cascades is crucial for various biological pathways, under the control of a range of scaffolding proteins. The BNIP-2 and Cdc42GAP Homology (BCH) domain is a highly conserved module that targets small GTPases and their regulators. Proteins bearing BCH domains are key for driving cell elongation, retraction, membrane protrusion, and other aspects of active morphogenesis during cell migration, myoblast differentiation, and neuritogenesis. We previously showed that the BCH domain of p50RhoGAP (ARHGAP1) sequesters RhoA from inactivation by its adjacent GAP domain; however, the underlying molecular mechanism for RhoA inactivation by p50RhoGAP remains unknown. Here, we report the crystal structure of the BCH domain of p50RhoGAP Schizosaccharomyces pombe and model the human p50RhoGAP BCH domain to understand its regulatory function using in vitro and cell line studies. We show that the BCH domain adopts an intertwined dimeric structure with asymmetric monomers and harbors a unique RhoA-binding loop and a lipid-binding pocket that anchors prenylated RhoA. Interestingly, the β5-strand of the BCH domain is involved in an intermolecular β-sheet, which is crucial for inhibition of the adjacent GAP domain. A destabilizing mutation in the β5-strand triggers the release of the GAP domain from autoinhibition. This renders p50RhoGAP active, thereby leading to RhoA inactivation and increased self-association of p50RhoGAP molecules via their BCH domains. Our results offer key insight into the concerted spatiotemporal regulation of Rho activity by BCH domain–containing proteins.

Small GTPases are molecular switches that cycle between an active GTP-bound state and an inactive GDP-bound state and are primarily involved in cytoskeletal reorganization during cell motility, morphogenesis, and cytokinesis (1, 2). These small GTPases are tightly controlled by activators and inactivators, such as guanine nucleotide exchange factors (GEFs) and GTPase-activating proteins (GAPs), respectively (3, 4), which are multidomain proteins that are themselves regulated through their interactions with other proteins, lipids, secondary messengers, and/or by posttranslational modifications (57). Despite our understanding of the mechanisms of action of GTPases, GAPs, and GEFs, little is known about how they are further regulated by other cellular proteins in tightly controlled local environments.The BNIP-2 and Cdc42GAP Homology (BCH) domain has emerged as a highly conserved and versatile scaffold protein domain that targets small GTPases, their GEFs, and GAPs to carry out various cellular processes in a spatial, temporal, and kinetic manner (815). BCH domain–containing proteins are classified into a distinct functional subclass of the CRAL_TRIO/Sec14 superfamily, with ∼175 BCH domain–containing proteins (in which 14 of them are in human) present across a range of eukaryotic species (16). Some well-studied BCH domain–containing proteins include BNIP-2, BNIP-H (CAYTAXIN), BNIP-XL, BNIP-Sα, p50RhoGAP (ARHGAP1), and BPGAP1 (ARHGAP8), with evidence to show their involvement in cell elongation, retraction, membrane protrusion, and other aspects of active morphogenesis during cell migration, growth activation and suppression, myoblast differentiation, and neuritogenesis (1721). Aside from interacting with small GTPases and their regulators, some of these proteins can also associate with other signaling proteins, such as fibroblast growth factor receptor tyrosine kinases, myogenic Cdo receptor, p38-MAP kinase, Mek2/MP1, and metabolic enzymes, such as glutaminase and ATP-citrate lyase (1726). Despite the functional diversity and versatility of BCH domain–containing proteins, the structure of the BCH domain and its various modes of interaction remain unknown. The BCH domain resembles the Sec14 domain (from the CRAL-TRIO family) (16, 27, 28), a domain with lipid-binding characteristics, which may suggest that the BCH domain could have a similar binding strategy. However, to date, the binding and the role of lipids in BCH domain function remain inconclusive.Of the BCH domain–containing proteins, we have focused on the structure and function of p50RhoGAP. p50RhoGAP comprises an N-terminal BCH domain and a C-terminal GAP domain separated by a proline-rich region. We found that p50RhoGAP contains a noncanonical RhoA-binding motif in its BCH domain and is associated with GAP-mediated cell rounding (13). Further, we showed previously that deletion of the BCH domain dramatically enhanced the activity of the adjacent GAP domain (13); however, the full dynamics of this interaction is unclear. Previously, it has been reported that the BCH and other domains regulate GAP activity in an autoinhibited manner (18, 21, 29, 30) involving the interactions of both the BCH and GAP domains, albeit the mechanism remains to be investigated. It has also been shown that a lipid moiety on Rac1 (a Rho GTPase) is necessary for its inactivation by p50RhoGAP (29, 31), which may imply a role in lipid binding. An understanding of how the BCH domain coordinates with the GAP domain to affect the local activity of RhoA and other GTPases would offer a previously unknown insight into the multifaceted regulation of Rho GTPase inactivation.To understand the BCH domain–mediated regulation of p50RhoGAP and RhoA activities, we have determined the crystal structure of a homologous p50RhoGAP BCH domain from S. pombe for functional interrogation. We show that the BCH domain adopts an intertwined dimeric structure with asymmetric monomers and harbors a unique RhoA-interacting loop and a lipid-binding pocket. Our results show that the lipid-binding region of the BCH domain helps to anchor the prenylation tail of RhoA while the loop interacts directly with RhoA. Moreover, we show that a mutation in the β5-strand releases the autoinhibition of the GAP domain by the BCH domain. This renders the GAP domain active, leading to RhoA inactivation and the associated phenotypic effects in yeast and HeLa cells. The released BCH domain also contributes to enhanced p50RhoGAP–p50RhoGAP interaction. Our findings offer crucial insights into the regulation of Rho signaling by BCH domain–containing proteins.  相似文献   

16.
Membraneless organelles containing the enzyme ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) are a common feature of organisms utilizing CO2 concentrating mechanisms to enhance photosynthetic carbon acquisition. In cyanobacteria and proteobacteria, the Rubisco condensate is encapsulated in a proteinaceous shell, collectively termed a carboxysome, while some algae and hornworts have evolved Rubisco condensates known as pyrenoids. In both cases, CO2 fixation is enhanced compared with the free enzyme. Previous mathematical models have attributed the improved function of carboxysomes to the generation of elevated CO2 within the organelle via a colocalized carbonic anhydrase (CA) and inwardly diffusing HCO3, which have accumulated in the cytoplasm via dedicated transporters. Here, we present a concept in which we consider the net of two protons produced in every Rubisco carboxylase reaction. We evaluate this in a reaction–diffusion compartment model to investigate functional advantages these protons may provide Rubisco condensates and carboxysomes, prior to the evolution of HCO3 accumulation. Our model highlights that diffusional resistance to reaction species within a condensate allows Rubisco-derived protons to drive the conversion of HCO3 to CO2 via colocalized CA, enhancing both condensate [CO2] and Rubisco rate. Protonation of Rubisco substrate (RuBP) and product (phosphoglycerate) plays an important role in modulating internal pH and CO2 generation. Application of the model to putative evolutionary ancestors, prior to contemporary cellular HCO3 accumulation, revealed photosynthetic enhancements along a logical sequence of advancements, via Rubisco condensation, to fully formed carboxysomes. Our model suggests that evolution of Rubisco condensation could be favored under low CO2 and low light environments.

Carbon dioxide (CO2) fixation into the biosphere has been primarily dependent upon action of the enzyme ribulose-1,5-bisphosphate carboxylase/oxygenase (Rubisco) over geological timescales. Rubisco is distinguished by the competitive inhibition of its carboxylation activity by the alternative substrate molecular oxygen (O2), leading to loss of CO2 and metabolic energy via a photorespiratory pathway in most phototrophs (1). Almost certainly the most abundant enzyme on the planet (2), Rubisco’s competing catalytic activities have required evolution of the enzyme, and/or its associated machinery, to maintain capture of sufficient carbon into organic molecules to drive life on Earth. In concert with geological weathering, the evolution of oxygenic photosynthesis ∼2.4 billion years ago has transformed the atmosphere from one rich in CO2 and low in O2 to one in which the relative abundances of these gases has overturned (3). Under these conditions, the Rubisco oxygenation reaction has increased, to the detriment of CO2 capture. This catalytic paradox has led to different adaptive solutions to ensure effective rates of photosynthetic CO2 fixation including the evolution of the kinetic properties of the enzyme (4), increases in Rubisco abundance in the leaves of many terrestrial C3 plants (5), and the evolution of diverse and complex CO2 concentrating mechanisms (CCMs) in many cyanobacteria, algae, and more recently hornworts, CAM, and C4 plants (6, 7).A defining characteristic of contemporary cyanobacteria is the encapsulation of their Rubisco enzymes within specialized, protein-encased microcompartments called carboxysomes (8). These microbodies are central to the cyanobacterial CCM, in which cellular bicarbonate (HCO3) is elevated by a combination of membrane-associated HCO3 pumps and CO2-to-HCO3 converting complexes (911), to drive CO2 production within the carboxysome by an internal carbonic anhydrase (CA; 12, 13). This process results in enhanced CO2 fixation, with a concomitant decrease in oxygenation, and is a proposed evolutionary adaptation to a low CO2 atmosphere (14, 15).An analogous CCM operates in many algal and hornwort species, which contain chloroplastic Rubisco condensates called pyrenoids (16, 17). Pyrenoids are liquid–liquid phase separated Rubisco aggregates, which lack the protein shell of a carboxysome (18). These CCMs accumulate HCO3 and convert it to CO2 within the pyrenoid to maximize CO2 fixation. Common to cyanobacterial and algal systems is the presence of unique Rubisco-binding proteins, enabling condensation of Rubisco from the bulk cytoplasm (1825). Condensation of proteins to form aggregates within the cell is increasingly recognized as a means by which cellular processes can be segregated and organized, across a broad range of biological systems (2628). The commonality of pyrenoid and carboxysome function (29) despite their disparate evolutionary histories (6), suggests a convergence of function driven by Rubisco condensation. In addition, dependency of functional CCMs on their pyrenoids or carboxysomes (30, 31) has led to the speculation that the evolution of Rubisco organization into membraneless organelles likely preceded systems which enabled elevated cellular HCO3 (14), raising the possibility that Rubisco condensation and encapsulation may have been the first steps in modern aquatic CCM evolution.We consider here that, in a primordial model system without active HCO3 accumulation, co-condensation of Rubisco and CA enzymes is beneficial for the elevation of internal CO2 because Rubisco carboxylation produces a net of two protons for every reaction turnover (SI Appendix, Fig. S1; 32, 33). These protons can be used within the condensate to convert HCO3 to CO2, with pH lowered and CO2 elevated as a result of restricted outward diffusion due to the high concentration of protein in the condensate and surrounding cell matrix acting as a barrier to diffusion. We propose that proton release within a primordial Rubisco condensate enabled the evolution of carboxysomes with enhanced carboxylation rates, prior to advancements which enabled cellular HCO3 accumulation.  相似文献   

17.
We previously described a new osteogenic growth factor, osteolectin/Clec11a, which is required for the maintenance of skeletal bone mass during adulthood. Osteolectin binds to Integrin α11 (Itga11), promoting Wnt pathway activation and osteogenic differentiation by leptin receptor+ (LepR+) stromal cells in the bone marrow. Parathyroid hormone (PTH) and sclerostin inhibitor (SOSTi) are bone anabolic agents that are administered to patients with osteoporosis. Here we tested whether osteolectin mediates the effects of PTH or SOSTi on bone formation. We discovered that PTH promoted Osteolectin expression by bone marrow stromal cells within hours of administration and that PTH treatment increased serum osteolectin levels in mice and humans. Osteolectin deficiency in mice attenuated Wnt pathway activation by PTH in bone marrow stromal cells and reduced the osteogenic response to PTH in vitro and in vivo. In contrast, SOSTi did not affect serum osteolectin levels and osteolectin was not required for SOSTi-induced bone formation. Combined administration of osteolectin and PTH, but not osteolectin and SOSTi, additively increased bone volume. PTH thus promotes osteolectin expression and osteolectin mediates part of the effect of PTH on bone formation.

The maintenance and repair of the skeleton require the generation of new bone cells throughout adult life. Osteoblasts are relatively short-lived cells that are constantly regenerated, partly by skeletal stem cells within the bone marrow (1). The main source of new osteoblasts in adult bone marrow is leptin receptor-expressing (LepR+) stromal cells (24). These cells include the multipotent skeletal stem cells that give rise to the fibroblast colony-forming cells (CFU-Fs) in the bone marrow (2), as well as restricted osteogenic progenitors (5) and adipocyte progenitors (68). LepR+ cells are a major source of osteoblasts for fracture repair (2) and growth factors for hematopoietic stem cell maintenance (911).One growth factor synthesized by LepR+ cells, as well as osteoblasts and osteocytes, is osteolectin/Clec11a, a secreted glycoprotein of the C-type lectin domain superfamily (5, 12, 13). Osteolectin is an osteogenic factor that promotes the maintenance of the adult skeleton by promoting the differentiation of LepR+ cells into osteoblasts. Osteolectin acts by binding to integrin α11β1, which is selectively expressed by LepR+ cells and osteoblasts, activating the Wnt pathway (12). Deficiency for either Osteolectin or Itga11 (the gene that encodes integrin α11) reduces osteogenesis during adulthood and causes early-onset osteoporosis in mice (12, 13). Recombinant osteolectin promotes osteogenic differentiation by bone marrow stromal cells in culture and daily injection of mice with osteolectin systemically promotes bone formation.Osteoporosis is a progressive condition characterized by reduced bone mass and increased fracture risk (14). Several factors contribute to osteoporosis development, including aging, estrogen insufficiency, mechanical unloading, and prolonged glucocorticoid use (14). Existing therapies include antiresorptive agents that slow bone loss, such as bisphosphonates (15, 16) and estrogens (17), and anabolic agents that increase bone formation, such as parathyroid hormone (PTH) (18), PTH-related protein (19), and sclerostin inhibitor (SOSTi) (20). While these therapies increase bone mass and reduce fracture risk, they are not a cure.PTH promotes both anabolic and catabolic bone remodeling (2124). PTH is synthesized by the parathyroid gland and regulates serum calcium levels, partly by regulating bone formation and bone resorption (2325). PTH1R is a PTH receptor (26, 27) that is strongly expressed by LepR+ bone marrow stromal cells (8, 2830). Recombinant human PTH (Teriparatide; amino acids 1 to 34) and synthetic PTH-related protein (Abaloparatide) are approved by the US Food and Drug Administration (FDA) for the treatment of osteoporosis (19, 31). Daily (intermittent) administration of PTH increases bone mass by promoting the differentiation of osteoblast progenitors, inhibiting osteoblast and osteocyte apoptosis, and reducing sclerostin levels (3235). PTH promotes osteoblast differentiation by activating Wnt and BMP signaling in bone marrow stromal cells (28, 36, 37), although the mechanisms by which it regulates Wnt pathway activation are complex and uncertain (38).Sclerostin is a secreted glycoprotein that inhibits Wnt pathway activation by binding to LRP5/6, a widely expressed Wnt receptor (7, 8), reducing bone formation (39, 40). Sclerostin is secreted by osteocytes (8, 41), negatively regulating bone formation by inhibiting the differentiation of osteoblasts (41, 42). SOSTi (Romosozumab) is a humanized monoclonal antibody that binds sclerostin, preventing binding to LRP5/6 and increasing Wnt pathway activation and bone formation (43). It is FDA-approved for the treatment of osteoporosis (20, 44) and has activity in rodents in addition to humans (45, 46).The discovery that osteolectin is a bone-forming growth factor raises the question of whether it mediates the effects of PTH or SOSTi on osteogenesis.  相似文献   

18.
19.
Enzymes that bear a nonnative or artificially introduced metal center can engender novel reactivity and enable new spectroscopic and structural studies. In the case of metal-organic cofactors, such as metalloporphyrins, no general methods exist to build and incorporate new-to-nature cofactor analogs in vivo. We report here that a common laboratory strain, Escherichia coli BL21(DE3), biosynthesizes cobalt protoporphyrin IX (CoPPIX) under iron-limited, cobalt-rich growth conditions. In supplemented minimal media containing CoCl2, the metabolically produced CoPPIX is directly incorporated into multiple hemoproteins in place of native heme b (FePPIX). Five cobalt-substituted proteins were successfully expressed with this new-to-nature cobalt porphyrin cofactor: myoglobin H64V V68A, dye decolorizing peroxidase, aldoxime dehydratase, cytochrome P450 119, and catalase. We show conclusively that these proteins incorporate CoPPIX, with the CoPPIX making up at least 95% of the total porphyrin content. In cases in which the native metal ligand is a sulfur or nitrogen, spectroscopic parameters are consistent with retention of native metal ligands. This method is an improvement on previous approaches with respect to both yield and ease-of-implementation. Significantly, this method overcomes a long-standing challenge to incorporate nonnatural cofactors through de novo biosynthesis. By utilizing a ubiquitous laboratory strain, this process will facilitate spectroscopic studies and the development of enzymes for CoPPIX-mediated biocatalysis.

The identity of a metal center often defines enzymatic activity, and swapping the native metal for an alternative one or introducing a new metal center has profound effects. More generally, the chemical utility of natural cofactors has inspired decades of study into synthetic analogs with distinct properties, and researchers have subsequently sought straightforward ways to put these novel cofactors back into proteins (1). Substituted metalloenzymes constitute one of the simplest cases. Changing the identity of the metal ion in metalloproteins has enabled powerful spectroscopic and functional studies of these proteins (210) in addition to new biocatalytic activities (1120). However, most methods for producing such proteins with new-to-nature cofactors are limited by the inability to produce the novel protein–cofactor complex in vivo.Hemoproteins, in particular, have been studied through metal substitution because of their important biological functions and utility as biocatalysts. Heme is a ubiquitous and versatile cofactor in biology, and heme-dependent proteins serve essential gas sensing functions (21), metabolize an array of xenobiotic molecules (22), and perform synthetically useful oxygen activation and radical-based chemistry (23). Metal-substituted hemoproteins have enabled key spectroscopic studies of hemoprotein function and the development of biocatalysts with novel reactivity. For example, electron paramagnetic resonance (EPR) studies on cobalt-substituted sperm whale myoglobin (CoMb) enabled detailed characterization of the paramagnetic CoMbO2 complex (3, 4, 24, 25). In analogous oxygen-binding studies in CoMb and cobalt-substituted hemoglobin (5, 6, 26), resonance Raman was used to identify the O–O stretching mode because cobalt-substituted proteins exhibit enhancement of this vibrational mode compared to the native iron proteins.Metal substitution has a profound effect on catalytic activity of hemoproteins, enabling numerous synthetic applications. Substitution of the native iron for cobalt in several hemoproteins, including a thermostable cytochrome c variant, enabled the reduction of water to H2 under aerobic, aqueous conditions (2729). Reconstitution of apoprotein with selected metalloporphyrins has been used to generate metal-substituted myoglobin and cytochrome P450s variants. These enzymes were effective as biocatalysts for C–H activation and carbene insertion reactions (1114). In a tour de force of directed evolution, which required purification and cofactor reconstitution of each individual variant, Hartwig and coworkers generated a cytochrome P450 variant that utilizes a nonnative Ir(Me)mesoporphyrin cofactor to perform desirable C–H activation chemistry (14). These activities may not be unique to the Ir-substituted protein, as synthetic cobalt porphyrin complexes have been shown to mediate a variety of Co(III)-aminyl and -alkyl radical transformations, including C–H activation (3032). Indeed, a number of cobalt porphyrin carbene complexes display significant carbon-centered radical character (3335), whereas the corresponding Fe-porphyrin complexes are closed shell species (36, 37), indicating that cobalt porphyrins may possess distinct, complementary modes of reactivity (3840).Inspired by these applications, researchers have sought strategies for generating metal-substituted hemoproteins. For many metalloproteins, metal substitution is carried out by removal of the native metal with a chelator and replacement with an alternate metal of similar coordination preference. This method is inapplicable to hemoproteins, as porphyrins do not readily exchange metal ions. Consequently, diverse methods have been employed to make metal-substituted hemoproteins (4146). Early on, copper, cobalt, nickel, and manganese-substituted horseradish peroxidase (HRP) were prepared by a multistep process that subjected protein to strong acid and organic solvents (41, 42). Variations of this method have been used repeatedly (24, 43, 4749). However, this method is applicable only to a narrow range of hemoproteins that tolerate the harsh treatment. With the advent of overexpression methods, significant improvement of metalloporphyrin-substituted protein yield was achieved by direct expression of the apoprotein and reconstitution with the desired metalloporphyrin in lysate prior to purification (50). Although this approach has many virtues, direct expression of apoprotein is ineffective for many hemoproteins, again limiting the utility of this method.As an alternative to the above in vitro approaches, researchers have pursued systems for direct in vivo expression of metal substituted hemoproteins. Two specialty strains of Escherichia coli (E. coli) were engineered to incorporate metalloporphyrin analogs from the growth medium into hemoproteins during protein expression. The engineered RP523 strain cannot biosynthesize heme and bears an uncharacterized heme permeability phenotype. Together, these two features enable this strain to assimilate and incorporate various metalloporphyrins into overexpressed hemoproteins with no background heme incorporation (44, 5153). However, heme auxotrophy makes RP523 cells exceedingly sensitive to O2, and, in many situations, RP523 cultures must be grown anaerobically. An alternative BL21(DE3)-based engineered strain harbors a plasmid bearing the heme transporter ChuA, which facilitates import of exogenous heme analogs (45). Production of metalloporphyrin-substituted protein with this ChuA-containing strain relies on growth in iron-limited minimal media, thereby diminishing heme biosynthesis. This method was used successfully to express metal-substituted versions of the heme domain of cytochrome P450 BM3 (45) and several myoglobin variants (11, 12). Because these cells biosynthesize a small quantity of their own heme, they are far more robust than the RP523 cells. Unfortunately, this advantage comes at the cost of increased heme contamination in the product protein (2 to 5%) (45).A set of intriguing papers reported the production of cobalt-substituted human cystathionine β-synthase (CoCBS) that relies on the de novo biosynthesis of CoPPIX from CoCl2 and δ-aminolevulinic acid (δALA), a biosynthetic precursor to heme (46, 54). This method yielded significant amounts of CoCBS—albeit with modest heme contamination (7.4%)—sufficient for spectroscopic and functional characterization of the CoPPIX-substituted protein (8, 46). As cobalt is known to be toxic to E. coli, the researchers passaged the CBS expression strain through cobalt-containing minimal media for 12 d, enabling the cells to adapt to high concentrations of cobalt prior to protein expression. It is plausible that this serial passaging alters the E. coli cells, enabling the biosynthesis of CoPPIX and in vivo production of metal-substituted protein. The adaptation process is slow (>10 d), and it is unknown how genomic instability under these mutagenic conditions affects the reproducibility of this passaging approach.The possibility of facile CoPPIX production is particularly attractive for future biocatalysis efforts. As described above, synthetic cobalt porphyrins have been shown to perform a range of radical-mediated reactions. The ability to produce a CoPPIX center in vivo may enable engineering these unusual reactivities via directed evolution in addition to spectroscopic applications. We therefore set out to explore the unusual phenotype of CoPPIX production by E. coli and to ascertain whether it was possible to efficiently biosynthesize cobalt-containing hemoproteins in vivo from a single “generalist” cell line. Our goal was to achieve an efficient and facile method of cobalt-substituted hemoprotein production with minimal contamination of the native cofactor. Herein, we report the surprising discovery that native E. coli BL21(DE3) can biosynthesize a new-to-nature CoPPIX cofactor (Fig. 1). We use this insight to produce cobalt-substituted hemoproteins in vivo without requirement for complex expression methods or specialized strains.Open in a separate windowFig. 1.Chemical structures of iron protoporphyrin IX (FePPIX or heme b), cobalt protoporphyrin IX (CoPPIX), and free base protoporphyrin IX (H2PPIX).  相似文献   

20.
Neuropeptides and neurotrophic factors secreted from dense core vesicles (DCVs) control many brain functions, but the calcium sensors that trigger their secretion remain unknown. Here, we show that in mouse hippocampal neurons, DCV fusion is strongly and equally reduced in synaptotagmin-1 (Syt1)- or Syt7-deficient neurons, but combined Syt1/Syt7 deficiency did not reduce fusion further. Cross-rescue, expression of Syt1 in Syt7-deficient neurons, or vice versa, completely restored fusion. Hence, both sensors are rate limiting, operating in a single pathway. Overexpression of either sensor in wild-type neurons confirmed this and increased fusion. Syt1 traveled with DCVs and was present on fusing DCVs, but Syt7 supported fusion largely from other locations. Finally, the duration of single DCV fusion events was reduced in Syt1-deficient but not Syt7-deficient neurons. In conclusion, two functionally redundant calcium sensors drive neuromodulator secretion in an expression-dependent manner. In addition, Syt1 has a unique role in regulating fusion pore duration.

To date, over 100 genes encoding neuropeptides and neurotrophic factors, together referred to as neuromodulators, are identified, and most neurons express neuromodulators and neuromodulator receptors (1). Neuromodulators travel through neurons in dense core vesicles (DCVs) and, upon secretion, regulate neuronal excitability, synaptic plasticity, and neurite outgrowth (24). Dysregulation of DCV secretion is linked to many brain disorders (57). However, the molecular mechanisms that regulate neuromodulator secretion remain largely elusive.Neuromodulator secretion, like neurotransmitter secretion from synaptic vesicles (SVs), is tightly controlled by Ca2+. The Ca2+ sensors that regulate secretion have been described for other secretory pathways but not for DCV exocytosis in neurons. Synaptotagmin (Syt) and Doc2a/b are good candidate sensors due to their interaction with SNARE complexes, phospholipids, and Ca2+ (811). The Syt family consists of 17 paralogs (12, 13). Eight show Ca2+-dependent lipid binding: Syt1 to 3, Syt5 to 7, and Syt9 and 10 (14, 15). Syt1 mediates synchronous SV fusion (8), consistent with its low Ca2+-dependent lipid affinity (15, 16) and fast Ca2+/membrane dissociation kinetics (16, 17). Syt1 is also required for the fast fusion in chromaffin cells (18) and fast striatal dopamine release (19). Synaptotagmin-7 (Syt7), in contrast, drives asynchronous SV fusion (20), in line with its a higher Ca2+ affinity (15) and slower dissociation kinetics (16). Syt7 is also a major calcium sensor for neuroendocrine secretion (21) and secretion in pancreatic cells (2224). Other sensors include Syt4, which negatively regulates brain-derived neurothropic factor (25) and oxytocin release (26), in line with its Ca2+ independency. Syt9 regulates hormone secretion in the anterior pituitary (27) and, together with Syt1, secretion from PC12 cells (28, 29). Syt10 controls growth factor secretion (30). However, Syt9 and Syt10 expression is highly restricted in the brain (3133). Hence, the calcium sensors for neuronal DCV fusion remain largely elusive. Because DCVs are generally not located close to Ca2+ channels (34), we hypothesized that DCV fusion is triggered by high-affinity Ca2+ sensors. Because of their important roles in vesicle secretion, their Ca2+ binding ability, and their high expression levels in the brain (20, 31, 3538), we addressed the roles of Doc2a/b, Syt1, and Syt7 in neuronal DCV fusion.In this study, we used primary Doc2a/b-, Syt1-, and Syt7-null (knockout, KO) neurons expressing DCV fusion reporters (34, 3941) with single-vesicle resolution. We show that both Syt1 and Syt7, but not Doc2a/b, are required for ∼60 to 90% of DCV fusion events. Deficiency of both Syt1 and Syt7 did not produce an additive effect, suggesting they function in the same pathway. Syt1 overexpression (Syt1-OE) rescued DCV fusion in Syt7-null neurons, and vice versa, indicating that the two proteins compensate for each other in DCV secretion. Moreover, overexpression of Syt1 or Syt7 in wild-type (WT) neurons increased DCV fusion, suggesting they are both rate limiting for DCV secretion. We conclude that DCV fusion requires two calcium sensors, Syt1 and Syt7, that act in a single/serial pathway and that both sensors regulate fusion in a rate-limiting and dose-dependent manner.  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