首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
3.
Biological membranes can dramatically accelerate the aggregation of normally soluble protein molecules into amyloid fibrils and alter the fibril morphologies, yet the molecular mechanisms through which this accelerated nucleation takes place are not yet understood. Here, we develop a coarse-grained model to systematically explore the effect that the structural properties of the lipid membrane and the nature of protein–membrane interactions have on the nucleation rates of amyloid fibrils. We identify two physically distinct nucleation pathways—protein-rich and lipid-rich—and quantify how the membrane fluidity and protein–membrane affinity control the relative importance of those molecular pathways. We find that the membrane’s susceptibility to reshaping and being incorporated into the fibrillar aggregates is a key determinant of its ability to promote protein aggregation. We then characterize the rates and the free-energy profile associated with this heterogeneous nucleation process, in which the surface itself participates in the aggregate structure. Finally, we compare quantitatively our data to experiments on membrane-catalyzed amyloid aggregation of α-synuclein, a protein implicated in Parkinson’s disease that predominately nucleates on membranes. More generally, our results provide a framework for understanding macromolecular aggregation on lipid membranes in a broad biological and biotechnological context.

The aggregation of normally soluble proteins into β-sheet-rich amyloid fibrils is a common form of protein assembly that has broad implications across biomedical and biotechnological sciences, in contexts as diverse as the molecular origins of neurodegenerative disorders to the production of functional materials (1, 2). The presence of surfaces and interfaces can strongly influence amyloid aggregation, either catalyzing or inhibiting it, depending on the nature of the surface. This effect has been studied for the cases of amyloid nucleation on nanoparticles (35), on flat surfaces (610), and on the surface of amyloid fibrils themselves (11, 12).Lipid bilayers are a unique type of surface, which is ubiquitous in biology and is the main contributor to the large surface-to-volume ratio characteristic of biological systems. They are highly dynamic, self-assembled structures that can induce structural changes in the proteins bound to them (13, 14) and markedly affect protein-aggregation propensities (15, 16). While nucleation on the surfaces of lipid membranes can influence fibril formation dramatically, alternative surfactant-driven fibrillation pathways in solution have been proposed at surfactant concentrations where formation of bilayer structures is not observed (17).Increasing experimental evidence supports the principle that the interaction between amyloidogenic proteins and the lipid cell membrane catalyzes in vivo amyloid nucleation, which is involved in debilitating pathologies. Remarkably, through surface-driven catalysis, lipid bilayers can enhance the kinetics of α-synuclein aggregation, the protein involved in Parkinson’s disease, by over three orders of magnitude with respect to nucleation in solution (18).Bilayer membranes can exist in different structural phases and can undergo local and global phase changes. A large body of work has focused on exploring how the membrane’s dynamical properties, such as its fluidity, relate to amyloid aggregation of bound proteins (1926).For instance, fluid membranes, constituted of short and saturated lipid chains, were found to most effectively catalyze the nucleation of α-synuclein (19), while less-fluid membranes composed of long lipid chains had less catalytic power. Furthermore, the addition of cholesterol to lipid membranes was found to alter its fluidity and govern the nucleation rate of Aβ42 (25), a peptide implicated in Alzheimer’s disease. In these cases, the physical properties of the membrane are controlled through variations in its composition, and decoupling the role of the membrane’s physical properties from its chemical specificity is extremely challenging.The question we focus on here is how the microscopic steps that drive amyloid nucleation at the membrane surface are altered by the inherently dynamic nature of lipid bilayers.Computer simulations can be of great help in this case, enabling us to systematically investigate the role of the physical and chemical properties of lipid membranes independently from one another, thus helping to identify key players behind membrane-driven amyloid nucleation.In this work, we develop a coarse-grained Monte Carlo model for studying the nucleation of amyloidogenic proteins on lipid membranes. We use it to identify the microscopic mechanisms which connect the membrane fluidity, the rate of amyloid nucleation, and the morphology of amyloid aggregates. We find that the membrane most efficiently catalyzes amyloid nucleation by donating its lipids to the nucleating fibril, which depends 1) on the lipid solubility and often correlates with membrane fluidity, and 2) the affinity of proteins to the membrane. This interdependence controls both the morphology of the resulting aggregates, which can range from protein-rich to lipid-rich, and the rate of fibril formation. We then discuss how our results provide a mechanistic explanation for a number of recent experimental observations regarding accelerated nucleation kinetics on fluid membranes (19), lipid–protein coaggregation (27), and altered aggregate morphology (23). Furthermore, the framework developed here offers a platform for studying strategies for bypassing amyloid nucleation in a cellular context.  相似文献   

4.
Engineered cytokines are gaining importance in cancer therapy, but these products are often limited by toxicity, especially at early time points after intravenous administration. 4-1BB is a member of the tumor necrosis factor receptor superfamily, which has been considered as a target for therapeutic strategies with agonistic antibodies or using its cognate cytokine ligand, 4-1BBL. Here we describe the engineering of an antibody fusion protein, termed F8-4-1BBL, that does not exhibit cytokine activity in solution but regains biological activity on antigen binding. F8-4-1BBL bound specifically to its cognate antigen, the alternatively spliced EDA domain of fibronectin, and selectively localized to tumors in vivo, as evidenced by quantitative biodistribution experiments. The product promoted a potent antitumor activity in various mouse models of cancer without apparent toxicity at the doses used. F8-4-1BBL represents a prototype for antibody-cytokine fusion proteins, which conditionally display “activity on demand” properties at the site of disease on antigen binding and reduce toxicity to normal tissues.

Cytokines are immunomodulatory proteins that have been considered for pharmaceutical applications in the treatment of cancer patients (13) and other types of disease (2). There is a growing interest in the use of engineered cytokine products as anticancer drugs, capable of boosting the action of T cells and natural killer (NK) cells against tumors (3, 4), alone or in combination with immune checkpoint inhibitors (3, 57).Recombinant cytokine products on the market include interleukin-2 (IL-2) (Proleukin) (8, 9), IL-11 (Neumega) (10, 11), tumor necrosis factor (TNF; Beromun) (12), interferon (IFN)-α (Roferon A, Intron A) (13, 14), IFN-β (Avonex, Rebif, Betaseron) (15, 16), IFN-γ (Actimmune) (17), granulocyte colony-stimulating factor (Neupogen) (18), and granulocyte macrophage colony-stimulating factor (Leukine) (19, 20). The recommended dose is typically very low (often <1 mg/d) (2123), as cytokines may exert biological activity in the subnanomolar concentration range (24). Various strategies have been proposed to develop cytokine products with improved therapeutic index. Protein PEGylation or Fc fusions may lead to prolonged circulation time in the bloodstream, allowing the administration of low doses of active payload (25, 26). In some implementations, cleavable polyethylene glycol polymers may be considered, yielding prodrugs that regain activity at later time points (27). Alternatively, tumor-homing antibody fusions have been developed, since the preferential concentration of cytokine payloads at the tumor site has been shown in preclinical models to potentiate therapeutic activity, helping spare normal tissues (2834). Various antibody-cytokine fusions are currently being investigated in clinical trials for the treatment of cancer and of chronic inflammatory conditions (reviewed in refs. 2, 33, 3537).Antibody-cytokine fusions display biological activity immediately after injection into patients, which may lead to unwanted toxicity and prevent escalation to therapeutically active dosage regimens (9, 22, 38). In the case of proinflammatory payloads (e.g., IL-2, IL-12, TNF-α), common side effects include hypotension, nausea, and vomiting, as well as flu-like symptoms (24, 3942). These side effects typically disappear when the cytokine concentration drops below a critical threshold, thus providing a rationale for slow-infusion administration procedures (43). It would be highly desirable to generate antibody-cytokine fusion proteins with excellent tumor-targeting properties and with “activity on demand”— biological activity that is conditionally gained on antigen binding at the site of disease, helping spare normal tissues.Here we describe a fusion protein consisting of the F8 antibody specific to the alternatively spliced extra domain A (EDA) of fibronectin (44, 45) and of murine 4-1BBL, which did not exhibit cytokine activity in solution but could regain potent biological activity on antigen binding. The antigen (EDA+ fibronectin) is conserved from mouse to man (46), is virtually undetectable in normal adult tissues (with the exception of the placenta, endometrium, and some vessels in the ovaries), but is expressed in the majority of human malignancies (44, 45, 47, 48). 4-1BBL, a member of the TNF superfamily (49), is expressed on antigen-presenting cells (50, 51) and binds to its receptor, 4-1BB, which is up-regulated on activated cytotoxic T cells (52), activated dendritic cells (52), activated NK and NKT cells (53), and regulatory T cells (54). Signaling through 4-1BB on cytotoxic T cells protects them from activation-induced cell death and skews the cells toward a more memory-like phenotype (55, 56).We engineered nine formats of the F8-4-1BBL fusion protein, one of which exhibited superior performance in quantitative biodistribution studies and conditional gain of cytokine activity on antigen binding. The antigen-dependent reconstitution of the biological activity of the immunostimulatory payload represents an example of an antibody fusion protein with “activity on demand.” The fusion protein was potently active against different types of cancer without apparent toxicity at the doses used. The EDA of fibronectin is a particularly attractive antigen for cancer therapy in view of its high selectivity, stability, and abundant expression in most tumor types (44, 45, 47, 48).  相似文献   

5.
Myristoylation is a posttranslational modification that plays diverse functional roles in many protein species. The myristate moiety is considered insufficient for protein–membrane associations unless additional membrane-affinity motifs, such as a stretch of positively charged residues, are present. Here, we report that the electrically neutral N-terminal fragment of the protein kinase A catalytic subunit (PKA-C), in which myristoylation is the only functional motif, is sufficient for membrane association. This myristoylation can associate a fraction of PKA-C molecules or fluorescent proteins (FPs) to the plasma membrane in neuronal dendrites. The net neutral charge of the PKA-C N terminus is evolutionally conserved, even though its membrane affinity can be readily tuned by changing charges near the myristoylation site. The observed membrane association, while moderate, is sufficient to concentrate PKA activity at the membrane by nearly 20-fold and is required for PKA regulation of AMPA receptors at neuronal synapses. Our results indicate that myristoylation may be sufficient to drive functionally significant membrane association in the absence of canonical assisting motifs. This provides a revised conceptual base for the understanding of how myristoylation regulates protein functions.

Myristoylation is a major type of posttranslational modification that occurs at the N terminus of a myriad of proteins (14). Depending on the target, myristoylation can contribute to the structure, stability, protein–protein interactions, and subcellular localization of the modified proteins (24). In particular, myristoylation often facilitates protein association with the membrane. However, it is thought that, with an acyl chain of only 14 carbons, myristate confers insufficient energy for stable association of a protein with the membrane (5, 6). Subsequent studies have shown that a second membrane-affinity motif, such as a stretch of basic residues or a second lipid modification, is required for the membrane association of several myristoylated proteins (reviewed in refs. 24). When the second membrane-affinity motif is removed or neutralized, either physiologically or via mutagenesis, the membrane localization of the protein is disrupted. Thus, the canonical view is that myristoylation alone is not sufficient to provide a functionally significant association of a protein with the plasma membrane, even though myristoylation has been observed to be associated with reconstituted lipid bilayers (7).Myristoylation was first discovered in the catalytic subunit of protein kinase A (PKA) (1, 8), which is a primary mediator of the second messenger cAMP that plays diverse essential roles in nearly all organisms, from bacteria to humans. At rest, PKA is a tetrameric protein that consists of two regulatory subunits (PKA-Rs) and two catalytic subunits (PKA-Cs). PKA holoenzymes are anchored to specific subcellular locations via the binding of PKA-R with A-Kinase anchoring proteins (912). In the presence of cAMP, PKA-C is released from PKA-R and becomes an active kinase (8, 1316).Despite being myristoylated, PKA-C is thought to function as a cytosolic protein because of its high solubility (14, 15). Consistently, a PKA-C mutant with disrupted myristoylation has been shown to support the phosphorylation of certain substrates and to maintain several PKA functions in heterologous cells (17). Structural studies found that the PKA-C myristoylation is folded into a hydrophobic pocket, and it was proposed that this myristoylation serves a structural role (1820). This view has started to shift based on recent reports showing that activated PKA-C can associate with the membrane in a myristoylation-dependent manner (16, 21, 22), including in neurons. However, the extent to which PKA-C associates with the plasma membrane in living cells and its functional significance are not known. Furthermore, as discussed above, myristolyation-mediated membrane association is thought to require a second-membrane motif. The identity of this second membrane-affinity motif has not been determined. Therefore, we set out to address these questions.  相似文献   

6.
Heterozygous point mutations of α-synuclein (α-syn) have been linked to the early onset and rapid progression of familial Parkinson’s diseases (fPD). However, the interplay between hereditary mutant and wild-type (WT) α-syn and its role in the exacerbated pathology of α-syn in fPD progression are poorly understood. Here, we find that WT mice inoculated with the human E46K mutant α-syn fibril (hE46K) strain develop early-onset motor deficit and morphologically different α-syn aggregation compared with those inoculated with the human WT fibril (hWT) strain. By using cryo-electron microscopy, we reveal at the near-atomic level that the hE46K strain induces both human and mouse WT α-syn monomers to form the fibril structure of the hE46K strain. Moreover, the induced hWT strain inherits most of the pathological traits of the hE46K strain as well. Our work suggests that the structural and pathological features of mutant strains could be propagated by the WT α-syn in such a way that the mutant pathology would be amplified in fPD.

α-Synuclein (α-Syn) is the main component of Lewy bodies, which serve as the common histological hallmark of Parkinson’s disease (PD) and other synucleinopathies (1, 2). α-Syn fibrillation and cell-to-cell transmission in the brain play essential roles in disease progression (35). Interestingly, WT α-syn could form fibrils with distinct polymorphs, which exhibit disparate seeding capability in vitro and induce distinct neuropathologies in mouse models (610). Therefore, it is proposed that α-syn fibril polymorphism may underlie clinicopathological variability of synucleinopathies (6, 9). In fPD, several single-point mutations of SNCA have been identified, which are linked to early-onset, severe, and highly heterogeneous clinical symptoms (1113). These mutations have been reported to influence either the physiological or pathological function of α-syn (14). For instance, A30P weakens while E46K strengthens α-syn membrane binding affinity that may affect its function in synaptic vesicle trafficking (14, 15). E46K, A53T, G51D, and H50Q have been found to alter the aggregation kinetics of α-syn in different manners (1517). Recently, several cryogenic electron microscopy (cryo-EM) studies revealed that α-syn with these mutations forms diverse fibril structures that are distinct from the WT α-syn fibrils (1826). Whether and how hereditary mutations induced fibril polymorphism contributes to the early-onset and exacerbated pathology in fPD remains to be elucidated. More importantly, most fPD patients are heterozygous for SNCA mutations (12, 13, 27, 28), which leads to another critical question: could mutant fibrils cross-seed WT α-syn to orchestrate neuropathology in fPD patients?E46K mutation is one of the eight disease-causing mutations on SNCA originally identified from a Spanish family with autosomal-dominant PD (11). E46K-associated fPD features early-onset motor symptoms and rapid progression of dementia with Lewy bodies (11). Studies have shown that E46K mutant has higher neurotoxicity than WT α-syn in neurons and mouse models overexpressing α-syn (2932). The underlying mechanism is debatable. Some reported that E46K promotes the formation of soluble species of α-syn without affecting the insoluble fraction (29, 30), while others suggested that E46K mutation may destabilize α-syn tetramer and induce aggregation (31, 32). Our previous study showed that E46K mutation disrupts the salt bridge between E46 and K80 in the WT fibril strain and rearranges α-syn into a different polymorph (33). Compared with the WT strain, the E46K fibril strain is prone to be fragmented due to its smaller and less stable fibril core (33). Intriguingly, the E46K strain exhibits higher seeding ability in vitro, suggesting that it might induce neuropathology different from the WT strain in vivo (33).In this study, we found that human E46K and WT fibril strains (referred to as hE46K and hWT strains) induced α-syn aggregates with distinct morphologies in mice. Mice injected with the hE46K strain developed more α-syn aggregation and early-onset motor deficits compared with the mice injected with the hWT strain. Notably, the hE46K strain was capable of cross-seeding both human and mouse WT (mWT) α-syn to form fibrils (named as hWTcs and mWTcs). The cross-seeded fibrils replicated the structure and seeding capability of the hE46K template both in vitro and in vivo. Our results suggest that the hE46K strain could propagate its structure as well as the seeding properties to the WT monomer so as to amplify the α-syn pathology in fPD.  相似文献   

7.
Protein aggregation into amyloid fibrils is associated with multiple neurodegenerative diseases, including Parkinson’s disease. Kinetic data and biophysical characterization have shown that the secondary nucleation pathway highly accelerates aggregation via the absorption of monomeric protein on the surface of amyloid fibrils. Here, we used NMR and electron paramagnetic resonance spectroscopy to investigate the interaction of monomeric α-synuclein (α-Syn) with its fibrillar form. We demonstrate that α-Syn monomers interact transiently via their positively charged N terminus with the negatively charged flexible C-terminal ends of the fibrils. These intermolecular interactions reduce intramolecular contacts in monomeric α-Syn, yielding further unfolding of the partially collapsed intrinsically disordered states of α-Syn along with a possible increase in the local concentration of soluble α-Syn and alignment of individual monomers on the fibril surface. Our data indicate that intramolecular unfolding critically contributes to the aggregation kinetics of α-Syn during secondary nucleation.

Synucleinopathies, including Parkinson’s disease (PD), are associated with the accumulation of intracellular neuronal aggregates termed as Lewy bodies and Lewy neuritis, which contain high concentration of the protein α-synuclein (α-Syn) in an aggregated state (1, 2). The disease-relevant role of α-Syn is further highlighted by mutations in the α-Syn gene (SNCA) causing familial PD [i.e., A30P (3), E46K (4), H50Q (5), G51D (6), A53E (7), and A53T (8)] and the duplication or triplication of the SNCA leading to early-onset PD in affected families (9, 10). α-Syn is a 140-residue intrinsically disordered protein (IDP) in solution (11) but adopts a helical structure in the presence of acidic lipid surfaces (12, 13). The positively charged N terminus (residues 1 to 60) is rich in lysine residues and contains KTKEGV binding repeats associated with vesicle binding (14). Moreover, the N-terminal domain includes all known SNCA familial PD mutations. The central region (residues 61 to 95) defines the non-amyloid-β component (NAC) (15), which is essential for α-Syn aggregation (16), while the C terminus (residues 96 to 140) is highly negatively charged.In vitro, α-Syn forms polymorphic amyloid fibrils (1719) with unique arrangements of cross-β-sheet motifs (2022). When injected into model animals, these fibrils induce a PD-like pathology (23) where the aggregation pathway of α-Syn plays a key role in the development of the disease (24). A detailed analysis of the aggregation kinetics of α-Syn into amyloids is therefore important toward understanding the toxic mechanisms relevant for synucleinopathies.Amyloid formation of α-Syn is very sensitive to solution conditions, including pH (25), temperature (26), and salt concentration (27). It further requires the presence of an air–water interface (28) or negatively charged lipid membranes (29) for which α-Syn has a high affinity. Previous studies suggest that amyloid fibril growth of α-Syn occurs via a nucleation-dependent polymerization reaction (30). Following a fairly slow primary nucleus formation, α-Syn fibrils are elongated by addition of single monomers. In a next step, the amyloid fibrils multiply by fragmentation or can catalyze the formation of new amyloids from monomers on their surface—a process known as secondary nucleation that was first described for sickle cell anemia 40 y ago (31). Fragmentation and secondary nucleation critically depend on the fibril mass and accelerate the aggregation kinetics (30). In the case of α-Syn aggregation under quiescent condition fragmentation does not exist and only the described secondary nucleation process occurs. While detailed kinetic experiments showed no significant secondary nucleation at pH 7, it strongly contributes at pH values lower than 6 (25, 30). However, mechanistic or structural information of the secondary nucleation process in α-Syn aggregation has been lacking so far.In this study we investigated the structural properties of α-Syn monomer–fibril interactions by NMR and electron paramagnetic resonance (EPR) spectroscopy. Our results provide insights into how monomeric α-Syn transiently interacts in vitro via its positively charged N terminus with the negatively charged C-terminal residues of the α-Syn fibrils, giving detailed insights into the mechanism of the secondary nucleation process.  相似文献   

8.
Membrane bending is a ubiquitous cellular process that is required for membrane traffic, cell motility, organelle biogenesis, and cell division. Proteins that bind to membranes using specific structural features, such as wedge-like amphipathic helices and crescent-shaped scaffolds, are thought to be the primary drivers of membrane bending. However, many membrane-binding proteins have substantial regions of intrinsic disorder which lack a stable three-dimensional structure. Interestingly, many of these disordered domains have recently been found to form networks stabilized by weak, multivalent contacts, leading to assembly of protein liquid phases on membrane surfaces. Here we ask how membrane-associated protein liquids impact membrane curvature. We find that protein phase separation on the surfaces of synthetic and cell-derived membrane vesicles creates a substantial compressive stress in the plane of the membrane. This stress drives the membrane to bend inward, creating protein-lined membrane tubules. A simple mechanical model of this process accurately predicts the experimentally measured relationship between the rigidity of the membrane and the diameter of the membrane tubules. Discovery of this mechanism, which may be relevant to a broad range of cellular protrusions, illustrates that membrane remodeling is not exclusive to structured scaffolds but can also be driven by the rapidly emerging class of liquid-like protein networks that assemble at membranes.

From endocytic buds (1) to needle-like filopodial protrusions (2), curved membrane surfaces play critical roles in many cellular processes (3). The energetic cost of creating these highly curved surfaces is considerable, such that spontaneous membrane fluctuations are insufficient to establish and stabilize the shapes of cellular membranes (4). Instead, work during the past two decades has revealed that interactions between proteins and lipids drive membrane curvature (5). Multiple physical mechanisms underlie the ability of proteins to shape membrane surfaces. These include amphipathic helices that insert like wedges into one leaflet of the membrane, creating an interleaflet area mismatch that drives curvature (6). Alternatively, proteins with inherently curved membrane binding domains such as BAR domains, dynamin, and ESCRTs act as scaffolds that can stabilize curved membrane geometries (7, 8). While each of these mechanisms relies on structured protein domains, we have recently reported that intrinsically disordered proteins, which lack a stable three-dimensional structure, can also be potent drivers of membrane bending (9, 10). Specifically, when noninteracting disordered domains are crowded together in cellular structures, steric repulsion among them drives the membrane to buckle outward, taking on a curved shape.Interestingly, rather than repelling one another, many disordered proteins have recently been found to assemble together via weak, multivalent interactions, forming networks that have the physical properties of liquids (11). Notably, recent studies have suggested that liquid–liquid phase separation of membrane-bound proteins plays an important role in diverse cellular processes including nucleation of actin filaments (12), immunological signaling (13), and assembly of virions (14).How might liquid–liquid phase separation of proteins at membrane surfaces impact membrane curvature? To address this question, we examined phase separation of the N-terminal low-complexity domain of fused in sarcoma, FUS LC, on the surfaces of synthetic and cell-derived membrane vesicles. FUS LC was chosen as a model protein for this study because it is among the most thoroughly characterized examples of a domain that undergoes liquid–liquid protein phase separation in solution (15). Here, we assemble FUS LC on membrane surfaces using an N-terminal histidine tag (16) that binds strongly to lipids with Ni-NTA headgroups. As FUS LC accumulated at the membrane surface, we observed protein phase separation in the two-dimensional plane of the membrane followed by spontaneous inward bending of the membrane, such that protein-lined tubules were created. Similar tubules were observed with two other domains implicated in liquid–liquid phase separation, the low-complexity domain of hnRNPA2 (17) and the RGG domain of LAF-1 (18). Interestingly, the tubules had undulating morphologies, similar to a string of pearls. This phenomenon is associated with an area mismatch between the two membrane leaflets (19, 20), suggesting that protein phase separation pulls lipids toward one another, creating a net compressive stress on one side of the membrane. In line with this hypothesis, a continuum mechanics model, built on the standard Helfrich framework, recreated the tubule morphology when a compressive stress was imposed using spontaneous curvature on the outer membrane surface. Further, the model predicted that tubule diameter should increase with increasing membrane rigidity and increasing rigidity ratio, trends confirmed by our experiments. Collectively, these findings suggest that protein phase separation on membrane surfaces generates considerable stresses that can drive the spontaneous assembly of membrane buds and tubules with physiologically relevant dimensions.  相似文献   

9.
Fatty acids (FAs) are central cellular metabolites that contribute to lipid synthesis, and can be stored or harvested for metabolic energy. Dysregulation in FA processing and storage causes toxic FA accumulation or altered membrane compositions and contributes to metabolic and neurological disorders. Saturated lipids are particularly detrimental to cells, but how lipid saturation levels are maintained remains poorly understood. Here, we identify the cerebellar ataxia spinocerebellar ataxia, autosomal recessive 20 (SCAR20)-associated protein Snx14, an endoplasmic reticulum (ER)–lipid droplet (LD) tethering protein, as a factor required to maintain the lipid saturation balance of cell membranes. We show that following saturated FA (SFA) treatment, the ER integrity of SNX14KO cells is compromised, and both SNX14KO cells and SCAR20 disease patient-derived cells are hypersensitive to SFA-mediated lipotoxic cell death. Using APEX2-based proximity labeling, we reveal the protein composition of Snx14-associated ER–LD contacts and define a functional interaction between Snx14 and Δ-9 FA desaturase SCD1. Lipidomic profiling reveals that SNX14KO cells increase membrane lipid saturation following exposure to palmitate, phenocopying cells with perturbed SCD1 activity. In line with this, SNX14KO cells manifest delayed FA processing and lipotoxicity, which can be rescued by SCD1 overexpression. Altogether, these mechanistic insights reveal a role for Snx14 in FA and ER homeostasis, defects in which may underlie the neuropathology of SCAR20.

Cells regularly internalize exogenous fatty acids (FAs) and must remodel their metabolic pathways to process and properly store FA loads. As it is a central cellular currency that can be stored, incorporated into membrane lipids, or harvested for energy, cells must balance FA uptake, processing, and oxidation to maintain homeostasis. Defects in any of these elevates intracellular free FAs (FFAs), which can act as detergents and damage organelles. Excessive membrane lipid saturation can also alter organelle function and contribute to cellular pathology, known as lipotoxicity (1, 2). Failure to properly maintain lipid compositions and storage contributes to many metabolic disorders (3), including type 2 diabetes (4), obesity (5), cardiac failure (6, 7), and various neurological diseases (8).Properties of FAs such as their degree of saturation and chain length are key determinants of their fate within the cell (9). High concentrations of saturated FAs (SFAs) in particular are highly toxic, as their incorporation into organelles affects membrane fluidity and can trigger lipotoxicity and cell death (1013). To prevent this, cells desaturate SFAs into monounsaturated FAs (MUFAs) before they are subsequently incorporated into membrane glycerophospholipids or stored as triglycerides (TGs) in lipid droplets (LDs). LD production provides a lipid reservoir to sequester otherwise toxic FAs, providing a metabolic buffer to maintain lipid homeostasis (14, 15).As LDs are created by and emerge from the ER network, interorganelle communication between the ER and LDs is vital for LD biogenesis (16). Consequently, numerous proteins that contribute to LD biogenesis, such as seipin (17, 18) and the diacylglyceride acyltransferase (DGAT) (19), are implicated in ER–LD cross-talk. Previously, we identified Snx14, a sorting nexin (SNX) protein linked to the cerebellar ataxia disease spinocerebellar ataxia, autosomal recessive 20 (SCAR20) (2022), as a novel factor that promotes FA-stimulated LD growth at ER–LD contacts (23, 24). Snx14 is an ER-anchored integral membrane protein. During periods of elevated FA flux, Snx14 is recruited to ER–LD contact sites, where it promotes the incorporation of FAs into TG as LDs grow (23). In line with this, SNX14KO cells exhibit defective LD morphology following oleate addition, implying Snx14 is required for proper FA storage in LDs. Related studies of Snx14 homologs in yeast and Drosophila indicate a conserved role for Snx14-family proteins in FA homeostasis and LD biogenesis (25, 26).Despite these insights, why humans with Snx14 loss-of-function mutations develop the cerebellar ataxia disease SCAR20 remains enigmatic. Given the proposed role of Snx14 in lipid metabolism, and that numerous neurological pathologies arise through defects in ER lipid homeostasis (2729), here we investigated whether Snx14 loss alters the ability of cells to maintain lipid homeostasis in response to FA influx. Our findings indicate that Snx14-deficient cells are hypersensitive to SFA exposure and manifest defects in ER morphology and ER-associated lipid metabolism.  相似文献   

10.
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   

11.
Interactions between proteins lie at the heart of numerous biological processes and are essential for the proper functioning of the cell. Although the importance of hydrophobic residues in driving protein interactions is universally accepted, a characterization of protein hydrophobicity, which informs its interactions, has remained elusive. The challenge lies in capturing the collective response of the protein hydration waters to the nanoscale chemical and topographical protein patterns, which determine protein hydrophobicity. To address this challenge, here, we employ specialized molecular simulations wherein water molecules are systematically displaced from the protein hydration shell; by identifying protein regions that relinquish their waters more readily than others, we are then able to uncover the most hydrophobic protein patches. Surprisingly, such patches contain a large fraction of polar/charged atoms and have chemical compositions that are similar to the more hydrophilic protein patches. Importantly, we also find a striking correspondence between the most hydrophobic protein patches and regions that mediate protein interactions. Our work thus establishes a computational framework for characterizing the emergent hydrophobicity of amphiphilic solutes, such as proteins, which display nanoscale heterogeneity, and for uncovering their interaction interfaces.

Protein–protein interactions play a crucial role in numerous biological processes, ranging from signal transduction and immune response to protein aggregation and phase behavior (13). Consequently, being able to understand, predict, and modulate protein interactions has important implications for understanding cellular processes and mitigating the progression of disease (4, 5). A necessary first step toward this ambitious goal is uncovering the interfaces through which proteins interact (68). In principle, identifying hydrophobic protein regions, which interact weakly with water, should be a promising strategy for uncovering protein interaction interfaces (9, 10). Indeed, the release of weakly interacting hydration waters from hydrophobic regions can drive protein interactions, as well as other aqueous assemblies (1113). However, even when the structure of a protein is available at atomistic resolution, it is challenging to identify its hydrophobic patches because they are not uniformly nonpolar, but display variations in polarity and charge at the nanoscale. Moreover, the emergent hydrophobicity of a protein patch stems from the collective response of protein hydration waters to the nanoscale chemical and topographical patterns displayed by the patch (1420) and cannot be captured by simply counting the number of nonpolar groups in the patch, or even through more involved additive approaches, such as hydropathy scales or surface-area models (2128).To address this challenge, we build upon seminal theoretical advances and molecular simulation studies, which have shown that near a hydrophobic surface, it is easier to disrupt surface–water interactions and form interfacial cavities (2934). To uncover protein regions that have the weakest interactions with water, here, we employ specialized molecular simulations, wherein protein–water interactions are disrupted by systematically displacing water molecules from the protein hydration shell (3537). By identifying the protein patches that nucleate cavities most readily in our simulations, we are then able to uncover the most hydrophobic protein regions. Interestingly, we find that both hydrophobic and hydrophilic protein patches are highly heterogeneous and contain comparable numbers of nonpolar and polar atoms. Our results thus highlight the nontrivial relationship between the chemical composition of protein patches and their emergent hydrophobicity (2426), and further emphasize the importance of accounting for the collective solvent response in characterizing protein hydrophobicity (16). We then interrogate whether the most hydrophobic protein patches, which nucleate cavities readily, are also likely to mediate protein interactions. Across five proteins that participate in either homodimer or heterodimer formation, we find that roughly 60 to 70% of interfacial contacts and only about 10 to 20% of noncontacts nucleate cavities. Our work thus provides a versatile computational framework for characterizing hydrophobicity and uncovering interaction interfaces of not just proteins, but also of other complex amphiphilic solutes, such as cavitands, dendrimers, and patchy nanoparticles (3841).  相似文献   

12.
Photosynthetic species evolved to protect their light-harvesting apparatus from photoxidative damage driven by intracellular redox conditions or environmental conditions. The Fenna–Matthews–Olson (FMO) pigment–protein complex from green sulfur bacteria exhibits redox-dependent quenching behavior partially due to two internal cysteine residues. Here, we show evidence that a photosynthetic complex exploits the quantum mechanics of vibronic mixing to activate an oxidative photoprotective mechanism. We use two-dimensional electronic spectroscopy (2DES) to capture energy transfer dynamics in wild-type and cysteine-deficient FMO mutant proteins under both reducing and oxidizing conditions. Under reducing conditions, we find equal energy transfer through the exciton 4–1 and 4–2-1 pathways because the exciton 4–1 energy gap is vibronically coupled with a bacteriochlorophyll-a vibrational mode. Under oxidizing conditions, however, the resonance of the exciton 4–1 energy gap is detuned from the vibrational mode, causing excitons to preferentially steer through the indirect 4–2-1 pathway to increase the likelihood of exciton quenching. We use a Redfield model to show that the complex achieves this effect by tuning the site III energy via the redox state of its internal cysteine residues. This result shows how pigment–protein complexes exploit the quantum mechanics of vibronic coupling to steer energy transfer.

Photosynthetic organisms convert solar photons into chemical energy by taking advantage of the quantum mechanical nature of their molecular systems and the chemistry of their environment (14). Antenna complexes, composed of one or more pigment–protein complexes, facilitate the first steps in the photosynthesis process: They absorb photons and determine which proportion of excitations to move to reaction centers, where charge separation occurs (4). In oxic environments, excitations can generate highly reactive singlet oxygen species. These pigment–protein complexes can quench excess excitations in these environments with molecular moieties such as quinones and cysteine residues (1, 57).The Fenna–Matthews–Olson (FMO) complex, a trimer of pigment–protein complexes found in the green sulfur bacterium Chlorobaculum tepidum (8), has emerged as a model system to study the photophysical properties of photosynthetic antenna complexes (919). Each subunit in the FMO complex contains eight bacteriochlorophyll-a site molecules (Protein Data Bank, ID code: 3ENI) that are coupled to form a basis of eight partially delocalized excited states called excitons (Fig. 1) (2023). Previous experiments on FMO have observed the presence of long-lived coherences in nonlinear spectroscopic signals at both cryogenic and physiological temperatures (11, 13). The coherent signals are thought to arise from some combination of electronic (2426), vibrational (1618), and vibronic (27) coherences in the system (2830). One previous study reported that the coherent signals in FMO remain unchanged upon mutagenesis of the protein, suggesting that the signals are ground state vibrational coherences (17). Others discuss the role of vibronic coupling, where electronic and nuclear degrees of freedom become coupled (29). Other dimeric model systems have demonstrated the regimes in which these vibronically coupled states produce coherent or incoherent transport and vibronic coherences (3133). Recent spectroscopic data has suggested that vibronic coupling plays a role in driving efficient energy transfer through photosynthetic complexes (27, 31, 33, 34), but to date there is no direct experimental evidence suggesting that biological systems use vibronic coupling as part of their biological function.Open in a separate windowFig. 1.(Left) Numbered sites and sidechains of cysteines C353 and C49 in the FMO pigment–protein complex (PDB ID code: 3ENI) (20). (Right) Site densities for excitons 4, 2, and 1 in reducing conditions with the energy transfer branching ratios for the WT oxidized and reduced protein. The saturation of pigments in each exciton denotes the relative contribution number to the exciton. The C353 residue is located near excitons 4 and 2, which have most electron density along one side of the complex, and other redox-active residues such as the Trp/Tyr chain. C353 and C49 surround site III, which contains the majority of exciton 1 density. Excitons 2 and 4 are generally delocalized over sites IV, V, and VII.It has been shown that redox conditions affect excited state properties in pigment-protein complexes, yet little is known about the underlying microscopic mechanisms for these effects (1, 9). Many commonly studied light-harvesting complexes—including the FMO complex (20), light-harvesting complex 2 (LH2) (35), the PC645 phycobiliprotein (36), and the cyanobacterial antenna complex isiA (37)—contain redox-active cysteine residues in close proximity to their chromophores. As the natural low light environment of C. tepidum does not necessitate photoprotective responses to light quantity and quality, its primary photoprotective mechanism concerns its response to oxidative stress. C. tepidum is an obligate anaerobe, but the presence of many active anoxygenic genes such as sodB for superoxide dismutase and roo for rubredoxin oxygen oxidoreductase (38) suggests that it is frequently exposed to molecular oxygen (7, 39). Using time-resolved fluorescence measurements, Orf et al. demonstrated that two cysteine residues in the FMO complex, C49 and C353, quench excitons under oxidizing conditions (1), which could protect the excitation from generating reactive oxygen species (7, 4042). In two-dimensional electronic spectroscopy (2DES) experiments, Allodi et al. showed that redox conditions in both the wild-type and C49A/C353A double-mutant proteins affect the ultrafast dynamics through the FMO complex (9, 43). The recent discovery that many proteins across the evolutionary landscape possess chains of tryptophan and tyrosine residues provides evidence that these redox-active residues may link the internal protein behavior with the chemistry of the surrounding environment (41, 43).In this paper, we present data showing that pigment–protein complexes tune the vibronic coupling of their chromophores and that the absence of this vibronic coupling activates an oxidative photoprotective mechanism. We use 2DES to show that a pair of cysteine residues in FMO, C49 and C353, can steer excitations toward quenching sites in oxic environments. The measured reaction rate constants demonstrate unusual nonmonotonic behavior. We then use a Redfield model to determine how the exciton energy transfer (EET) time constants arise from changing chlorophyll site energies and their system-bath couplings (44, 45). The analysis reveals that the cysteine residues tune the resonance between exciton 4–1 energy gap and an intramolecular chlorophyll vibration in reducing conditions to induce vibronic coupling and detune the resonance in oxidizing conditions. This redox-dependent modulation of the vibronic coupling steers excitations through different pathways in the complex to change the likelihood that they interact with exciton quenchers.  相似文献   

13.
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   

14.
Fusion-associated small transmembrane (FAST) proteins are a diverse family of nonstructural viral proteins. Once expressed on the plasma membrane of infected cells, they drive fusion with neighboring cells, increasing viral spread and pathogenicity. Unlike viral fusogens with tall ectodomains that pull two membranes together through conformational changes, FAST proteins have short fusogenic ectodomains that cannot bridge the intermembrane gap between neighboring cells. One orthoreovirus FAST protein, p14, has been shown to hijack the actin cytoskeleton to drive cell-cell fusion, but the actin adaptor-binding motif identified in p14 is not found in any other FAST protein. Here, we report that an evolutionarily divergent FAST protein, p22 from aquareovirus, also hijacks the actin cytoskeleton but does so through different adaptor proteins, Intersectin-1 and Cdc42, that trigger N-WASP–mediated branched actin assembly. We show that despite using different pathways, the cytoplasmic tail of p22 can replace that of p14 to create a potent chimeric fusogen, suggesting they are modular and play similar functional roles. When we directly couple p22 with the parallel filament nucleator formin instead of the branched actin nucleation promoting factor N-WASP, its ability to drive fusion is maintained, suggesting that localized mechanical pressure on the plasma membrane coupled to a membrane-disruptive ectodomain is sufficient to drive cell-cell fusion. This work points to a common biophysical strategy used by FAST proteins to push rather than pull membranes together to drive fusion, one that may be harnessed by other short fusogens responsible for physiological cell-cell fusion.

Aquareovirus and orthoreovirus are two genera of the Reoviridae family of segmented double-stranded RNA viruses that form multinucleated syncytia after infection, which can increase viral spread and pathogenicity (14). To drive cell-cell fusion, both aquareovirus and orthoreovirus express a nonstructural, fusion-associated small transmembrane (FAST) protein on the plasma membrane of infected cells. The FAST protein is not required for viral entry, and expression of FAST protein alone is sufficient to cause cells to fuse with naïve neighboring cells, forming large multinucleated syncytium (1, 2, 512), confirming they are bona fide cell-cell fusogens. Although they have similar function and topology in the membrane, FAST proteins from aquareovirus and orthoreovirus share minimal sequence identity (13). Based on phylogenetic analysis, they are hypothesized to have evolved from a common, likely nonfusogenic, ancestor 510 million years ago (4, 13, 14). Separate gain-of-function events are believed to have produced fusogenic proteins in both aquareovirus and orthoreovirus, with further divergence or acquisition events resulting in the diversity of FAST proteins found in reoviruses today (13).Aquareovirus and orthoreovirus FAST proteins are single-pass membrane proteins of fewer than 200 residues comprised of a mostly disordered cytoplasmic tail, a transmembrane domain, and a small ectodomain of fewer than 40 residues (1, 2). The membrane-disruptive ectodomains of FAST proteins typically have solvent-exposed hydrophobic residues and/or myristoylation motifs that are necessary for cell-cell fusion (5, 1517). In contrast to other cell-cell fusogens that fuse membranes by pulling them together using conformational changes in their ∼10 nm-tall ectodomains, the ectodomains of FAST proteins have minimal predicted secondary structure, are unlikely to undergo conformational changes to drive membrane fusion (1, 2), and extend only ∼1 nm above the bilayer (5, 18). How such short fusogens can overcome the ∼2 nm repulsive hydration barrier and larger barrier presented by cell surface proteins to reach and fuse with an opposing membrane (5, 18) has been a long-standing question for FAST proteins and other short cell-cell fusogens, such as myomixer and myomaker that are involved in myoblast fusion (1922).Recently, we found that the FAST protein from reptilian orthoreovirus, p14, hijacks the host cell actin cytoskeleton to drive cell-cell fusion by forming localized branched actin networks (23). This is accomplished through a c-src phosphorylated tyrosine motif, YVNI, in p14’s disordered cytoplasmic tail that binds to a host adaptor protein, Grb2, which then binds to N-WASP and nucleates branched actin assembly. We hypothesize that this directly couples local actin-generated forces to push p14’s short, fusogenic ectodomain into the opposing cell’s plasma membrane (23). While all FAST family proteins have similarly short ectodomains, it is unclear if this is a general strategy used by other FAST proteins to drive cell-cell fusion.Here, we report that a FAST protein from the divergent aquareovirus, p22, also hijacks the host actin cytoskeleton but does so using a molecular strategy distinct from that of the orthoreovirus FAST protein p14. Instead of binding to Grb2, we find that p22 binds to Intersectin-1 through an SH3 binding motif in its cytoplasmic tail, which binds Cdc42 to activate N-WASP–mediated branched actin assembly. We show that despite minimal sequence identity, the p22 cytoplasmic tail can be functionally swapped with that of p14, suggesting that while the cytoplasmic tails of the two FAST proteins evolved independently, they serve a similar function. By directly coupling the ectodomain to a different actin nucleator, we suggest that actin’s functional role is applying mechanical pressure to a fusogenic ectodomain at the plasma membrane. This biophysical role may be shared across other members of the FAST protein family and could be more generally employed by other cell-cell fusogens.  相似文献   

15.
Active matter comprises individually driven units that convert locally stored energy into mechanical motion. Interactions between driven units lead to a variety of nonequilibrium collective phenomena in active matter. One of such phenomena is anomalously large density fluctuations, which have been observed in both experiments and theories. Here we show that, on the contrary, density fluctuations in active matter can also be greatly suppressed. Our experiments are carried out with marine algae (Effreniumvoratum), which swim in circles at the air–liquid interfaces with two different eukaryotic flagella. Cell swimming generates fluid flow that leads to effective repulsions between cells in the far field. The long-range nature of such repulsive interactions suppresses density fluctuations and generates disordered hyperuniform states under a wide range of density conditions. Emergence of hyperuniformity and associated scaling exponent are quantitatively reproduced in a numerical model whose main ingredients are effective hydrodynamic interactions and uncorrelated random cell motion. Our results demonstrate the existence of disordered hyperuniform states in active matter and suggest the possibility of using hydrodynamic flow for self-assembly in active matter.

Active matter exists over a wide range of spatial and temporal scales (16) from animal groups (7, 8) to robot swarms (911), to cell colonies and tissues (1216), to cytoskeletal extracts (1720), to man-made microswimmers (2125). Constituent particles in active matter systems are driven out of thermal equilibrium at the individual level; they interact to develop a wealth of intriguing collective phenomena, including clustering (13, 22, 24), flocking (11, 26), swarming (12, 13), spontaneous flow (14, 20), and giant density fluctuations (10, 11). Many of these observed phenomena have been successfully described by particle-based or continuum models (16), which highlight the important roles of both individual motility and interparticle interactions in determining system dynamics.Current active matter research focuses primarily on linearly swimming particles which have a symmetric body and self-propel along one of the symmetry axes. However, a perfect alignment between the propulsion direction and body axis is rarely found in reality. Deviation from such a perfect alignment leads to a persistent curvature in the microswimmer trajectories; examples of such circle microswimmers include anisotropic artificial micromotors (27, 28), self-propelled nematic droplets (29, 30), magnetotactic bacteria and Janus particles in rotating external fields (31, 32), Janus particle in viscoelastic medium (33), and sperm and bacteria near interfaces (34, 35). Chiral motility of circle microswimmers, as predicted by theoretical and numerical investigations, can lead to a range of interesting collective phenomena in circular microswimmers, including vortex structures (36, 37), localization in traps (38), enhanced flocking (39), and hyperuniform states (40). However, experimental verifications of these predictions are limited (32, 35), a situation mainly due to the scarcity of suitable experimental systems.Here we address this challenge by investigating marine algae Effrenium voratum (41, 42). At air–liquid interfaces, E.voratum cells swim in circles via two eukaryotic flagella: a transverse flagellum encircling the cellular anteroposterior axis and a longitudinal one running posteriorly. Over a wide range of densities, circling E.voratum cells self-organize into disordered hyperuniform states with suppressed density fluctuations at large length scales. Hyperuniformity (43, 44) has been considered as a new form of material order which leads to novel functionalities (4549); it has been observed in many systems, including avian photoreceptor patterns (50), amorphous ices (51), amorphous silica (52), ultracold atoms (53), soft matter systems (5461), and stochastic models (6264). Our work demonstrates the existence of hyperuniformity in active matter and shows that hydrodynamic interactions can be used to construct hyperuniform states.  相似文献   

16.
Leukocyte homing driven by the chemokine CCL21 is pivotal for adaptive immunity because it controls dendritic cell (DC) and T cell migration through CCR7. ACKR4 scavenges CCL21 and has been shown to play an essential role in DC trafficking at the steady state and during immune responses to tumors and cutaneous inflammation. However, the mechanism by which ACKR4 regulates peripheral DC migration is unknown, and the extent to which it regulates CCL21 in steady-state skin and lymph nodes (LNs) is contested. Specifically, our previous findings that CCL21 levels are increased in LNs of ACKR4-deficient mice [I. Comerford et al., Blood 116, 4130–4140 (2010)] were refuted [M. H. Ulvmar et al., Nat. Immunol. 15, 623–630 (2014)], and no differences in CCL21 levels in steady-state skin of ACKR4-deficient mice were reported despite compromised CCR7-dependent DC egress in these animals [S. A. Bryce et al., J. Immunol. 196, 3341–3353 (2016)]. Here, we resolve these issues and reveal that two forms of CCL21, full-length immobilized and cleaved soluble CCL21, exist in steady-state barrier tissues, and both are regulated by ACKR4. Without ACKR4, extracellular CCL21 gradients in barrier sites are saturated and nonfunctional, DCs cannot home directly to lymphatic vessels, and excess soluble CCL21 from peripheral tissues pollutes downstream LNs. The results identify the mechanism by which ACKR4 controls DC migration in barrier tissues and reveal a complex mode of CCL21 regulation in vivo, which enhances understanding of functional chemokine gradient formation.

CCL21 is a chemokine that mediates recruitment of multiple leukocyte subsets through CCR7-mediated signaling during the steady state and inflammation. CCL21 plays crucial roles in priming adaptive immunity via governing egress of dendritic cells (DCs) from barrier tissues and T cell entry and positioning in secondary lymphoid organs (15). A well-characterized site of CCL21 gradient formation is the skin, where CCL21 is secreted by lymphatic endothelial cells (LECs) and immobilized on extracellular heparan sulfate moieties via interactions with the charged, elongated C-terminal tail of CCL21 (68). Here, immobilized CCL21 gradients are essential for interstitial DC trafficking toward lymphatic vessels (LVs) (8), after which CCL21 further contributes to LV attachment (9), infiltration (10), downstream luminal migration (11), and migration from the lymph node (LN) subcapsular sinus (SCS) to the paracortex (12). In-vitro studies have shown that the C-terminal tail of CCL21 can also be proteolytically cleaved by mature DCs to generate solubilized CCL21 (13), with signaling properties distinct from another soluble CCR7 ligand, CCL19 (14, 15). While CCL19 is dispensable for steady-state DC migration (16), important questions regarding the in vivo processing and function of cleaved CCL21 remain.Both forms of CCL21 are also ligands for the atypical chemokine receptor ACKR4 (17), which regulates chemokine bioavailability rather than directly mediating cell migration. ACKR4 expression has been identified in multiple barrier tissues (1820) and lymphoid tissues (12, 21) where expression is largely restricted to stromal cell populations, with the exception of germinal-center B cells (22). Despite clear evidence of ACKR4 scavenging of CCL21 in vitro, the extent to which it regulates CCL21 in vivo is disputed. We have shown increased CCL21 in the LNs of Ackr4−/− mice, which was associated with exacerbated Th17 responses in autoimmunity (23) and an ACKR4-dependent increase in CCL21 in tumors that promotes antitumor immunity (24). However, no differences in dermal CCL21 abundance were previously reported in steady-state Ackr4−/− mice despite steady-state CCR7-dependent DC migratory defects being independent of CCL19 (19), and the contribution of ACKR4 in regulating LN CCL21 abundance has been disputed despite a clear role for ACKR4 in maintaining interfollicular CCL21 gradients in LN (12). These discrepancies have remained unresolved but point to previously unrecognized complexity in ACKR4-dependent regulation of CCL21 in both barrier and lymphoid tissues.  相似文献   

17.
18.
Physiological functioning and homeostasis of the brain rely on finely tuned synaptic transmission, which involves nanoscale alignment between presynaptic neurotransmitter-release machinery and postsynaptic receptors. However, the molecular identity and physiological significance of transsynaptic nanoalignment remain incompletely understood. Here, we report that epilepsy gene products, a secreted protein LGI1 and its receptor ADAM22, govern transsynaptic nanoalignment to prevent epilepsy. We found that LGI1–ADAM22 instructs PSD-95 family membrane-associated guanylate kinases (MAGUKs) to organize transsynaptic protein networks, including NMDA/AMPA receptors, Kv1 channels, and LRRTM4–Neurexin adhesion molecules. Adam22ΔC5/ΔC5 knock-in mice devoid of the ADAM22–MAGUK interaction display lethal epilepsy of hippocampal origin, representing the mouse model for ADAM22-related epileptic encephalopathy. This model shows less-condensed PSD-95 nanodomains, disordered transsynaptic nanoalignment, and decreased excitatory synaptic transmission in the hippocampus. Strikingly, without ADAM22 binding, PSD-95 cannot potentiate AMPA receptor-mediated synaptic transmission. Furthermore, forced coexpression of ADAM22 and PSD-95 reconstitutes nano-condensates in nonneuronal cells. Collectively, this study reveals LGI1–ADAM22–MAGUK as an essential component of transsynaptic nanoarchitecture for precise synaptic transmission and epilepsy prevention.

Epilepsy, characterized by unprovoked, recurrent seizures, affects 1 to 2% of the population worldwide. Many genes that cause inherited epilepsy when mutated encode ion channels, and dysregulated synaptic transmission often causes epilepsy (1, 2). Although antiepileptic drugs have mainly targeted ion channels, they are not always effective and have adverse effects. It is therefore important to clarify the detailed processes for synaptic transmission and how they are affected in epilepsy.Recent superresolution imaging of the synapse reveals previously overlooked subsynaptic nano-organizations and pre- and postsynaptic nanodomains (36), and mathematical simulation suggests their nanometer-scale coordination in individual synapses for efficient synaptic transmission: presynaptic neurotransmitter release machinery and postsynaptic receptors precisely align across the synaptic cleft to make “transsynaptic nanocolumns” (7, 8).So far, numerous transsynaptic cell-adhesion molecules have been identified (912), including presynaptic Neurexins and type IIa receptor protein tyrosine phosphatases (PTPδ, PTPσ, and LAR) and postsynaptic Neuroligins, LRRTMs, NGL-3, IL1RAPL1, Slitrks, and SALMs. Neurexins–Neuroligins have attracted particular attention because of their synaptogenic activities when overexpressed and their genetic association with neuropsychiatric disorders (e.g., autism). Another type of transsynaptic adhesion complex mediated by synaptically secreted Cblns (e.g., Neurexin–Cbln1–GluD2) promotes synapse formation and maintenance (1315). Genetic studies in Caenorhabditis elegans show that secreted Ce-Punctin, the ortholog of the mammalian ADAMTS-like family, specifies cholinergic versus GABAergic identity of postsynaptic domains and functions as an extracellular synaptic organizer (16). However, the molecular identity and in vivo physiological significance of transsynaptic nanocolumns remain incompletely understood.LGI1, a neuronal secreted protein, and its receptor ADAM22 have recently emerged as major determinants of brain excitability (17) as 1) mutations in the LGI1 gene cause autosomal dominant lateral temporal lobe epilepsy (18); 2) mutations in the ADAM22 gene cause infantile epileptic encephalopathy with intractable seizures and intellectual disability (19, 20); 3) Lgi1 or Adam22 knockout mice display lethal epilepsy (2124); and 4) autoantibodies against LGI1 cause limbic encephalitis characterized by seizures and amnesia (2528). Functionally, LGI1–ADAM22 regulates AMPA receptor (AMPAR) and NMDA receptor (NMDAR)-mediated synaptic transmission (17, 22, 29) and Kv1 channel-mediated neuronal excitability (30, 31). Recent structural analysis shows that LGI1 and ADAM22 form a 2:2 heterotetrameric assembly (ADAM22–LGI1–LGI1–ADAM22) (32), suggesting the transsynaptic configuration.In this study, we identify ADAM22-mediated synaptic protein networks in the brain, including pre- and postsynaptic MAGUKs and their functional bindings to transmembrane proteins (NMDA/AMPA glutamate receptors, voltage-dependent ion channels, cell-adhesion molecules, and vesicle-fusion machinery). ADAM22 knock-in mice lacking the MAGUK-binding motif show lethal epilepsy of hippocampal origin. In this mouse, postsynaptic PSD-95 nano-assembly as well as nano-scale alignment between pre- and postsynaptic proteins are significantly impaired. Importantly, PSD-95 is no longer able to modulate AMPAR-mediated synaptic transmission without binding to ADAM22. These findings establish that LGI1–ADAM22 instructs MAGUKs to organize transsynaptic nanocolumns and guarantee the stable brain activity.  相似文献   

19.
The mechanism and pore architecture of the Tat complex during transport of folded substrates remain a mystery, partly due to rapid dissociation after translocation. In contrast, the proteinaceous SecY pore is a persistent structure that needs only to undergo conformational shifts between “closed” and “opened” states when translocating unfolded substrate chains. Where the proteinaceous pore model describes the SecY pore well, the toroidal pore model better accounts for the high-energy barrier that must be overcome when transporting a folded substrate through the hydrophobic bilayer in Tat transport. Membrane conductance behavior can, in principle, be used to distinguish between toroidal and proteinaceous pores, as illustrated in the examination of many antimicrobial peptides as well as mitochondrial Bax and Bid. Here, we measure the electrochromic shift (ECS) decay as a proxy for conductance in isolated thylakoids, both during protein transport and with constitutively assembled translocons. We find that membranes with the constitutively assembled Tat complex and those undergoing Tat transport display conductance characteristics similar to those of resting membranes. Membranes undergoing Sec transport and those with the substrate-engaged SecY pore result in significantly more rapid electric field decay. The responsiveness of the ECS signal in membranes with active SecY recalls the steep relationship between applied voltage and conductance in a proteinaceous pore, while the nonaccelerated electric field decay with both Tat transport and the constitutive Tat complex under the same electric field is consistent with the behavior of a toroidal pore.

The twin arginine translocation pathway is uniquely able to transport fully folded substrates in an ATP-independent manner, relying instead on an electrochemical gradient (i.e., the proton motive force, or pmf) across the transporting membrane. It is crucial to the transport of substrates requiring various cofactors and hetero-oligomeric complexes in prokaryotes and of substrates vital to photosynthetic machinery in thylakoids (1). In plant mitochondria, the Rieske Fe/S protein required for the biogenesis of complex III is transported by the Tat pathway (26). It is implicated in both the virulence and antibiotic resistance of various infectious bacteria (712) and has been studied for its potential in biotechnology applications (1315). The uniqueness of Tat functionality and its appearance across the kingdoms of life make it a valuable research target for crop modification, biotechnology, and pathogenesis. Unfortunately, much of the knowledge about its mechanism has been hard won, and the pore structure remains a mystery, likely due to the transient nature of the active complex.The active Tat complex in thylakoids consists of three core subunits, Tha4, Hcf106, and cpTatC, which are homologous to the bacterial TatA, TatB, and TatC, respectively (1, 16). An N-terminal signal peptide with a twin-arginine motif inserts into the cis-side leaflet at the TatBC receptor complex (1719). Subsequent oligomerization of TatA subunits (2022) at the TatBC receptor complex results in rings of varying sizes (22, 23), but it is unclear whether these structures represent transport pores. Of particular note is the short TatA transmembrane helix (TMH). Composed of only 16 residues, the solid-state NMR solution suggests that the TMH must tilt and draw a portion of the cis-side amphipathic helix (APH) into the membrane in order to cross the bilayer (24), establishing a resting state of hydrophobic mismatch. During transport, a conformational shift increasing the angle between the TMH and APH results in exacerbated hydrophobic mismatch, as the APH is moved radially away from the center and the TMH is pulled up toward the cis-side in the active state (25, 26). For both native and foreign substrates, the Tat-targeted signal peptide and the pmf are sufficient to cause assembly of the active translocon and achieve transport (2731). After the translocation event, the complex dissociates into TatA and TatBC components (1, 15, 16) with the exception of some residual TatA bound to the receptor complex in a nonactive state and a spectrum of smaller TatA oligomers (32).Within the thylakoid membrane, it is useful to compare the Tat complex with the general secretory translocon (Sec) because they both function in the same membrane environment (1, 33). Sec translocation first requires recruitment of the substrate to the soluble SecA ATPase to form the substrate–SecA complex, which is then recruited to the SecY pore (1). In the inactive state, the proteinaceous SecY pore prevents ion leakage through a combination of a trans-side plug domain and an internal array of hydrophobic residues (34). Following substrate–SecA docking, a conformational shift in SecY allows substrate movement through the open pore in an ATP-dependent process driven by SecA (35). In the mammalian homolog Sec61, leakage of NAD+ ions is recorded during ribosome-bound nascent chain transport in a fluorescence quenching study, suggesting the pore can reach 4 to 6 nm in diameter (36). However, X-ray structures of substrate-fused SecA complexed with SecY (35), conductance studies in ribosome-bound substrates engaged with SecY (37), and SecY plug deletion mutants (38) in Escherichia coli have estimated the open SecY pore diameter to range from 7.3 to 8.8 Å, almost 10-fold lower. This small diameter likely contributes to the restriction of ion movement during Sec transport (39).While the Sec machinery only transports unfolded substrates (40), the Tat pore accommodates substrates ranging from a single unfolded chain in an engineered system (13) to a folded substrate with an average diameter of 70 Å (41). This extended size range raises an interesting question about the pore architecture. In the Sec translocon, X-ray crystallography of the SecY channel in Methanocaldococcus jannaschii (42) and Thermotoga maritima (43) reveals that the SecY pore channel excludes lipids in both the resting state and when engaged with its SecA partner. Further structural work on the E. coli substrate-engaged SecA–SecY complex shows that the SecY channel excludes lipids during transport as well (35). No such structural information about the Tat pore exists, but functional data suggest that TatA plays an important role in the pore (20, 44, 45) and cryogenic electron microscopy structures of TatA oligomers reveal rings of an internal diameter ranging from 30 to 70 Å (23). During transport of the 17-kDa subunit of the oxygen-evolving complex (OE23), the Tat pathway exhibits very low ion leakage (46), estimated to be less than 1 pS. This is despite the exchange of 80,000 protons per substrate (47). Extensive mapping of subunit–subunit and subunit–substrate contacts has revealed no putative plug domain (20, 4851) that could be compared to that in the SecY protein.Pore architecture can be characterized by membrane conductance behavior. Conductance measured through proteinaceous pores representing the barrel-stave model has a very steep dependence on the voltage applied, whereas conductance in toroidal pores requires a larger voltage to be detected and increases more slowly in response to increasing voltage (5254). Performing similar experiments on the Tat and Sec translocons would require functional reconstitution of both complexes into an in vitro setting. However, decay of the electrochromic shift (ECS) signal can be used as a measure of ion conductance (46). A transient absorption peak at ∼515 nm arising from carotenoid pigments in response to the native electric field generated by charge separation in the photosynthetic reaction centers (55) can be measured by delivering a single-saturating flash. The decay rate of such a signal is a direct measurement of how quickly the electric field is dissipated by ion movement across the membrane.In the experiments reported herein, ECS signal decay rates revealing the conductance states of resting isolated membranes and those engaged in ongoing transport and in the presence of a constitutively assembled and/or substrate-engaged translocon are used to probe the pore architecture in the Tat and Sec complexes. Increased conductance across the thylakoid membrane is indicated by a more rapidly decaying ECS signal. We find that conductance in thylakoid membranes during Sec-mediated transport and substrate-engaged SecY is consistently higher than that during Tat-mediated transport and with the constitutively assembled Tat complex, respectively, despite the much larger Tat pore required to transport a fully folded substrate. This points to a difference not only in mechanism but in pore architecture between the two. Conductance behavior of membranes undergoing Sec-mediated transport is consistent with that of a proteinaceous pore, while the resistance demonstrated by membranes undergoing Tat-mediated transport is more reminiscent of toroidal pores.  相似文献   

20.
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   

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