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

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

2.
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   

3.
Autophagy is a catabolic pathway that provides self-nourishment and maintenance of cellular homeostasis. Autophagy is a fundamental cell protection pathway through metabolic recycling of various intracellular cargos and supplying the breakdown products. Here, we report an autophagy function in governing cell protection during cellular response to energy crisis through cell metabolic rewiring. We observe a role of selective type of autophagy in direct activation of cyclic AMP protein kinase A (PKA) and rejuvenation of mitochondrial function. Mechanistically, autophagy selectively degrades the inhibitory subunit RI of PKA holoenzyme through A-kinase–anchoring protein (AKAP) 11. AKAP11 acts as an autophagy receptor that recruits RI to autophagosomes via LC3. Glucose starvation induces AKAP11-dependent degradation of RI, resulting in PKA activation that potentiates PKA-cAMP response element-binding signaling, mitochondria respiration, and ATP production in accordance with mitochondrial elongation. AKAP11 deficiency inhibits PKA activation and impairs cell survival upon glucose starvation. Our results thus expand the view of autophagy cytoprotection mechanism by demonstrating selective autophagy in RI degradation and PKA activation that fuels the mitochondrial metabolism and confers cell resistance to glucose deprivation implicated in tumor growth.

Macroautophagy (henceforth autophagy) is a catabolic process that degrades various cellular cargos through lysosomes. The autophagy process includes the formation and trafficking of autophagosomes, which sequester the cellular cargos destined for the clearance. Autophagy is activated in response to nutrient deprivation or cellular injuries and serves as a recycling mechanism that maintains cellular homeostasis through degradation of cytoplasmic components. Autophagy provides cell self-nourishment and supports cellular metabolism by supplying breakdown products (1, 2); therefore, autophagy is a fundamental cell protection mechanism. Whether autophagy has a direct function beyond recycling of the breakdown molecules to maintain metabolic homeostasis and cell survival, however, is poorly understood. In certain cancer types, autophagy plays an important role in sustaining the aggressive growth of the tumor cells by enhancing cell metabolism. Although our group and others have previously shown an inhibitory function of Beclin 1–mediated autophagy in tumorigenesis (3, 4), the current view is that tumors, once established, rely heavily on autophagy to survive due to high metabolic demand. One potential mechanism is that the metabolic products generated by autophagy provide tumor cells with metabolic rewiring that enables them to survive even under nutrient-poor conditions (5, 6). However, it remains unclear whether autophagy plays a role beyond the production of metabolic fuel sources to maintain metabolic plasticity and tumor cell growth.Available evidence has demonstrated the selectivity of autophagy in the digestion of certain cellular cargoes mediated by autophagy adaptors/receptors. Characterization of the autophagy adaptors has shed light on the versatile physiological function of autophagy in the maintenance of the homeostasis for large molecules and cellular organelles (7, 8). These adaptors recognize and recruit selective cargos to autophagy machinery for degradation through direct interaction with yeast autophagy gene Atg8 homologs of mammalian LC3/GABARAP/Gate16 proteins. While a few autophagy receptors have been reported, it is clear that many more are yet to be identified (7, 8).The best-known signaling pathways that control the metabolic stress-induced autophagy are mediated by mTOR and AMPK kinases, both of which are the master regulators for cellular metabolism (9, 10). Cyclic adenosine monophosphate (cAMP)-dependent protein kinase (PKA) is also a key kinase of cell metabolism that governs diverse cellular pathways, including cellular glucose metabolism and bioenergetic processes (1113). Surprisingly, whether and how cAMP/PKA regulates autophagy or vice versa is poorly understood in mammals (1416). cAMP/PKA signaling has emerged over recent years as a key regulator for mitochondrial functions, highlighting the mechanism of cAMP/PKA in cellular metabolism control (17, 18). Despite an established role for PKA in the regulation of mitochondrial metabolism, whether autophagy and PKA converge to regulate metabolic reprogramming and cell survival remains unknown.The PKA holoenzyme consists of two regulatory subunits (R) and two catalytic subunits (C). The R subunits are inhibitory of catalytic kinase activity; upon binding to cAMP, the R subunit dissociates from C subunit, resulting in activation of PKA (19). Furthermore, the specific cellular functions of PKA are controlled by a number of A-kinase–anchoring proteins (AKAPs). The AKAPs bind the R subunits and restrict the PKA holoenzyme to various intracellular compartments, providing spatiotemporal regulation of PKA activity (20). However, the functions of many AKAPs are poorly characterized.Here, we report a role of autophagy that controls cellular metabolism beyond the production of metabolic sources—it activates cAMP/PKA kinase activity by selective degradation of the inhibitory subunit of R1α through autophagy receptor AKAP11 in response to glucose starvation. AKAP11-mediated cAMP/PKA activation leads to elevation of mitochondrial metabolism and cell protection. Our study reveals a previously unrecognized function of autophagy in metabolic rewiring of cells that promote cell survival under energy crisis. Our study thus suggests that selective autophagy induced RI degradation and PKA activation may contribute to the resistance of tumor cells to metabolic stress.  相似文献   

4.
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   

5.
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   

6.
A chromosome 1q21.3 region that is frequently amplified in diverse cancer types encodes phosphatidylinositol (PI)-4 kinase IIIβ (PI4KIIIβ), a key regulator of secretory vesicle biogenesis and trafficking. Chromosome 1q21.3–amplified lung adenocarcinoma (1q-LUAD) cells rely on PI4KIIIβ for Golgi-resident PI-4-phosphate (PI4P) synthesis, prosurvival effector protein secretion, and cell viability. Here, we show that 1q-LUAD cells subjected to prolonged PI4KIIIβ antagonist treatment acquire tolerance by activating an miR-218-5p–dependent competing endogenous RNA network that up-regulates PI4KIIα, which provides an alternative source of Golgi-resident PI4P that maintains prosurvival effector protein secretion and cell viability. These findings demonstrate an addiction to Golgi-resident PI4P synthesis in a genetically defined subset of cancers.

The term “oncogene addiction” was coined to describe cancer cells’ exquisite dependence on individual oncogenes to sustain the malignant phenotype (1, 2). Examples include the BCR-ABL oncogene produced by a chromosome 9:22 translocation in chronic myelogenous leukemia and the somatically mutated EGFR oncogene in lung adenocarcinoma (LUAD) (35). In both cancer types, the mutant kinases are bona fide oncogenes in vitro and in vivo (6, 7). Although patients treated with selective kinase inhibitors attain profound clinical responses (8), chronic exposure of patients to targeted therapeutics is followed by disease relapse owing to an almost universal reactivation of mutant kinase activity, demonstrating that most cancers retain an underlying addiction to oncogene-induced signaling pathways (2, 9). Elucidating the molecular underpinnings of oncogene reactivation may lead to improved therapeutic strategies.Heightened secretion of protumorigenic effector proteins promotes metastasis and acquired resistance to targeted therapeutics (10, 11). The conventional secretory pathway directs the transport of secretory vesicles from the endoplasmic reticulum to the plasma membrane via the Golgi apparatus (12). Tensile forces exerted on Golgi membranes activate secretory vesicle biogenesis and are mediated by a Golgi phosphoprotein-3 (GOLPH3)/F-actin protein complex (13). GOLPH3 binds to phosphatidylinositol (PI)-4-phosphate (PI4P), which tethers GOLPH3 to Golgi membranes and is generated by the Golgi-resident PI-4 kinases PI4KIIα and PI4KIIIβ (10, 13, 14).A chromosome 1q21.3 region that is frequently amplified in diverse cancer types encodes PI4KIIIβ (11, 15). Chromosome 1q21.3–amplified LUAD (1q-LUAD) cells undergo apoptosis following treatment with small-molecule PI4KIIIβ antagonists or depletion of PI4KIIIβ–dependent secreted proteins (11), establishing that PI4KIIIβ–dependent secreted proteins are prosurvival effectors in 1q-LUAD cells. On the basis of this conceptual framework, here, we postulated that chromosome 1q21.3 amplifications confer an addiction to Golgi-resident PI-4 kinase activity.  相似文献   

7.
8.
Binding of the spike protein of SARS-CoV-2 to the human angiotensin-converting enzyme 2 (ACE2) receptor triggers translocation of the virus into cells. Both the ACE2 receptor and the spike protein are heavily glycosylated, including at sites near their binding interface. We built fully glycosylated models of the ACE2 receptor bound to the receptor binding domain (RBD) of the SARS-CoV-2 spike protein. Using atomistic molecular dynamics (MD) simulations, we found that the glycosylation of the human ACE2 receptor contributes substantially to the binding of the virus. Interestingly, the glycans at two glycosylation sites, N90 and N322, have opposite effects on spike protein binding. The glycan at the N90 site partly covers the binding interface of the spike RBD. Therefore, this glycan can interfere with the binding of the spike protein and protect against docking of the virus to the cell. By contrast, the glycan at the N322 site interacts tightly with the RBD of the ACE2-bound spike protein and strengthens the complex. Remarkably, the N322 glycan binds to a conserved region of the spike protein identified previously as a cryptic epitope for a neutralizing antibody. By mapping the glycan binding sites, our MD simulations aid in the targeted development of neutralizing antibodies and SARS-CoV-2 fusion inhibitors.

Angiotensin-converting enzyme 2 (ACE2) is an enzyme that catalyzes the hydrolysis of angiotensin II into angiotensin (17) to counterbalance the ACE receptor in blood pressure control (1). A single transmembrane helix anchors ACE2 into the plasma membrane of cells in the lungs, arteries, heart, kidney, and intestines (2). The vasodilatory effect of ACE2 has made it a promising target for drugs treating cardiovascular diseases (3).ACE2 also serves as the entry point for several coronaviruses into cells, including SARS-CoV and SARS-CoV-2 (46). The binding of the spike protein of SARS-CoV and SARS-CoV-2 to the peptidase domain (PD) of ACE2 triggers endocytosis and translocation of both the virus and the ACE2 receptor into endosomes within cells (4). The human transmembrane serine protease 2, TMPRSS2, primes spike for efficient cell entry by cleaving its backbone at the boundary between the S1 and S2 subunits or within the S2 subunit (4). The structure of the ACE2 receptor in complex with the SARS-CoV-2 spike receptor binding domain (RBD) (79) reveals the major RBD interaction regions as helix H1 (Q24–Q42), a loop in a beta sheet (K353–R357), and the end of helix H2 (L79–Y83). With a 4-Å heavy-atom distance cutoff, 20 residues of ACE2 interact with 17 residues of the RBD, forming a buried interface of ∼1,700 Å2 (7).The structure of full-length ACE2 has been resolved in complex with B0AT1 (also known as SLC6A19) (9). B0AT1 is a sodium-dependent neutral amino acid transporter (10). ACE2 functions as chaperone for B0AT1 and is responsible for its trafficking to the plasma membrane of kidney and intestine epithelial cells (11). Although it was speculated that B0AT1 prevents ACE2 cleavage by TMPRSS2 and thus could suppress SARS-CoV-2 infection (9, 12), other studies showed that SARS-CoV-2 can infect human small intestinal enterocytes where ACE2 is expected to be in complex with B0AT1 (13).Both the ACE2 receptor and the spike protein are heavily glycosylated. Several glycosylation sites are near the binding interface (7, 9, 14, 15). Whereas the focus has largely been on amino acid interactions in the ACE2–spike binding interface (16, 17), the role of glycosylation in binding has been recognized (7, 1820). The extracellular domain of the ACE2 receptor has seven N-glycosylation sites (N53, N90, N103, N322, N432, N546, and N690) and several O-glycosylation sites (e.g., T730) (9, 14). Among ACE2 glycosylation sites, the only well-characterized position regarding the effect on the spike binding and viral infectivity is N90. It is known from earlier SARS-CoV studies that glycosylation at the N90 position might interfere with virus binding and infectivity (21). Also, recent genetic and biochemical studies showed that mutations of N90, which remove the glycosylation site directly, or of T92, which remove the glycosylation site indirectly by eliminating the glycosylation motif (NXT), increase the susceptibility to SARS-CoV-2 infection (22, 23).We use extensive molecular dynamics (MD) simulations to gain a detailed molecular-level understanding of how ACE2 glycosylation impacts the host–virus interactions. Glycosylation sites N90 and N322 of human ACE2 emerge as major determinants of its binding to SARS-CoV-2 spike. Remarkably, glycans at these sites have opposite effects, interfering with spike binding in one case, and strengthening binding in the other. Our findings provide direct guidance for the design of targeted antibodies and therapeutic inhibitors of viral entry.  相似文献   

9.
Lipid droplets (LDs) are intracellular organelles responsible for lipid storage, and they emerge from the endoplasmic reticulum (ER) upon the accumulation of neutral lipids, mostly triglycerides (TG), between the two leaflets of the ER membrane. LD biogenesis takes place at ER sites that are marked by the protein seipin, which subsequently recruits additional proteins to catalyze LD formation. Deletion of seipin, however, does not abolish LD biogenesis, and its precise role in controlling LD assembly remains unclear. Here, we use molecular dynamics simulations to investigate the molecular mechanism through which seipin promotes LD formation. We find that seipin clusters TG, as well as its precursor diacylglycerol, inside its unconventional ring-like oligomeric structure and that both its luminal and transmembrane regions contribute to this process. This mechanism is abolished upon mutations of polar residues involved in protein–TG interactions into hydrophobic residues. Our results suggest that seipin remodels the membrane of specific ER sites to prime them for LD biogenesis.

Lipid droplets (LDs) are the intracellular organelles responsible for fat accumulation (1). As such, they play a central role in lipid and cellular metabolism (14), and they are crucially involved in metabolic diseases such as lipodystrophy and obesity (57).Formation of LDs occurs in the endoplasmic reticulum (ER), where neutral lipids (NLs), namely triglycerides (TG) and cholesteryl esters, constituting the core of LDs are synthesized by acyltransferases that are essential for LD formation (8). The current model of LD formation posits that NLs are stored between the two leaflets of the ER bilayer, where they aggregate in nascent oblate lens-like structures with diameters of 40 to 60 nm (9) before complete maturation and budding toward the cytosol (1013).Recent experiments suggest that LDs form at specific ER sites marked by the protein seipin (14) upon arrival of its interaction partner protein promethin/LDAF1 (lipid droplet organization [LDO] in yeast) (1519). These recent observations confirm previous works showing that seipin, in addition to modulating LD budding and growth (14, 1921) and LD–ER contacts (22, 23), is also a major player in the early stages of LD formation, as deletion of seipin leads to TG accumulation in the ER and a delay in the formation of, possibly aberrant, LDs (20, 24).The role of seipin in LD formation is potentially coupled to its function in regulating lipid metabolism (25, 26) and notably that of phosphatidic acid (PA) (2731). Recently, seipin-positive ER loci have been shown to be part of a larger protein machinery that also includes membrane and lipid remodeling proteins of the TG synthesis pathway (32), most notably, Lipin (Pah1 in yeast) and FIT proteins (Yft2 and Scs3 in yeast), for which PA is either a known substrate (Lipin/Pah1) (33) or a likely one (FIT/Yft2/Scs3) (34).Despite this thorough characterization of the cellular role of seipin in LD formation, the molecular details of its mechanism remain mostly unclear. Recently, the structure of the luminal part of the seipin oligomer has been solved at 3.7 to 4.0 Å resolution using electron microscopy (27, 35), paving the way for the investigation of the relationship between its three-dimensional structure and its mode of action. These studies revealed that the luminal domain of seipin consists of an eight-stranded beta sandwich, together with a hydrophobic helix (HH), positioned toward the ER bilayer. Notably, the seipin oligomer assembles into a ring-like architecture, an unconventional assembly in lipid bilayers that rather resembles the shape of microbial pore-forming assemblies (36) or GroEL-GroES chaperones (37, 38).From a stochiometric point of view, both fluorescence and electron microscopy data are consistent with the presence of a single seipin oligomer per nascent LD (14, 15). Hence, the structure of the luminal part of seipin is consistent with two proposed modes of action: seipin could mark the sites of LD formation by controlling TG flow in and out of the nascent droplet (14), or, alternatively, seipin could help recognize and stabilize preexisting nascent droplets in the ER membrane (20, 21, 39). In both cases, however, the relationship between the role of seipin in LD formation and its ability to regulate lipid metabolism remains unclear.Here, we use coarse-grain (CG) molecular dynamics (MD) simulations to investigate the mechanism of seipin in molecular detail. We find that seipin is able to cluster TG molecules inside its ring-like structure and that both the transmembrane (TM) helices and the luminal domain contribute to this process. Diacylglycerol (DG), the lipid intermediate between TG and PA in the Kennedy pathway, also accumulates around seipin, further promoting the accumulation of TG at very low TG-to-phospholipids ratios. Our data suggest that by accumulating DG and TG molecules, seipin generates ER sites with a specific lipid composition that in turn could promote the sequential recruitment of additional TG- and DG-sensing proteins involved in LD formation, including promethin/LDOs, FIT/Yft2/Scs3, and perilipins.  相似文献   

10.
Staphylococcus aureus colonizes the skin of the majority of patients with atopic dermatitis (AD), and its presence increases disease severity. Adhesion of S. aureus to corneocytes in the stratum corneum is a key initial event in colonization, but the bacterial and host factors contributing to this process have not been defined. Here, we show that S. aureus interacts with the host protein corneodesmosin. Corneodesmosin is aberrantly displayed on the tips of villus-like projections that occur on the surface of AD corneocytes as a result of low levels of skin humectants known as natural moisturizing factor (NMF). An S. aureus mutant deficient in fibronectin binding protein B (FnBPB) and clumping factor B (ClfB) did not bind to corneodesmosin in vitro. Using surface plasmon resonance, we found that FnBPB and ClfB proteins bound with similar affinities. The S. aureus binding site was localized to the N-terminal glycine–serine-rich region of corneodesmosin. Atomic force microscopy showed that the N-terminal region was present on corneocytes containing low levels of NMF and that blocking it with an antibody inhibited binding of individual S. aureus cells to corneocytes. Finally, we found that S. aureus mutants deficient in FnBPB or ClfB have a reduced ability to adhere to low-NMF corneocytes from patients. In summary, we show that FnBPB and ClfB interact with the accessible N-terminal region of corneodesmosin on AD corneocytes, allowing S. aureus to take advantage of the aberrant display of corneodesmosin that accompanies low NMF in AD. This interaction facilitates the characteristic strong binding of S. aureus to AD corneocytes.

Atopic dermatitis (AD) is a chronic inflammatory skin disorder, affecting 15 to 20% of children (1, 2). During disease flares, patients experience painful inflamed skin lesions accompanied by intense itch. Epidermal barrier dysfunction, increased type 2 immune responses, and recurrent skin infections are features of AD (3). The most common cause of infection is Staphylococcus aureus. This bacterium colonizes the skin of the majority of AD patients (4, 5). Isolates representing several S. aureus lineages are recovered from AD skin; however, strains from the clonal complex 1 (CC1) lineage are the most frequently isolated (69). The burden of S. aureus on lesional and nonlesional skin correlates with severity of the disease (10, 11). S. aureus directly influences pathogenesis, and several factors produced by the bacterium increase inflammation and exacerbate AD symptoms, including staphylococcal superantigen B and delta-toxin (1215).Despite the clear association between S. aureus colonization and AD disease severity (11), the bacterial and host factor determinants underlying colonization are poorly understood (16). Adhesion is a critical early step in the colonization process. S. aureus adheres to corneocytes in the stratum corneum of AD skin (6, 17, 18). We previously found that clumping factor B (ClfB), a cell wall-anchored protein displayed on the surface of S. aureus, can mediate adhesion to corneocytes from AD patients (6). ClfB also binds to the alpha chain of fibrinogen and to the cornified envelope proteins loricrin and cytokeratin 10 (K10) in desquamated nasal epithelial cells (1921). To date, ClfB is the only bacterial factor known to promote adherence to corneocytes in AD. However, a ClfB-deficient mutant retained the ability to bind to corneocytes (6), suggesting that additional bacterial factors are at play.Filaggrin deficiency is common in patients with established AD and is either genetic or caused by down-regulation of gene expression by Th-2–type cytokines (2224). Filaggrin deficiency causes epidermal barrier defects and a loss of the hygroscopic filaggrin breakdown products that normally contribute to the natural moisturizing factor (NMF) in corneocytes (25). NMF comprises a collection of humectants, including filaggrin breakdown products urocanic acid and pyrrolidone acid, along with urea, citrate, lactate acid, and sugars, and is responsible for regulating hydration in the skin (26). Low-NMF levels are associated with a loss of hydration, increased disease severity, and abnormal corneocyte morphology (27). We showed recently that S. aureus binds more strongly to low-NMF AD corneocytes than to corneocytes with normal levels of NMF (18).Corneocytes with low NMF have very different surface topography when compared with corneocytes with normal levels of NMF (27). Aberrant “villus-like” projections (VPs) protrude from the surface of low-NMF corneocytes (18, 27). The protein corneodesmosin (CDSN) is confined to the cell–cell junctions between corneocytes in healthy skin, where homophilic interactions between the CDSN proteins on adjacent cells facilitate cell–cell cohesion (28). In AD patients, however, CDSN decorates the tips of the VPs on low-NMF corneocytes (27).This study aimed to elucidate a key component of S. aureus colonization by identifying the molecular determinants of adherence to AD corneocytes. We recognized that the occurrence of VPs on low-NMF corneocytes presents a different colonization surface to the bacterium and postulated that the accessibility of CDSN on the tips of VPs could influence pathogen adherence. We show that S. aureus can interact with CDSN and identify the S. aureus proteins promoting adherence to this host protein. We use single-cell and single-molecule atomic force microscopy (AFM), surface plasmon resonance (SPR), and ex vivo bacterial adherence studies with patient corneocytes to characterize this interaction. This study expands the repertoire of ligands for S. aureus and, crucially, links bacterial interactions with a host protein (CDSN) to binding to corneocytes taken from patients. Thus, our findings provide insights into the adhesion process and develop our understanding of the mechanisms underlying colonization of the skin of AD patients by S. aureus.  相似文献   

11.
We previously described a new osteogenic growth factor, osteolectin/Clec11a, which is required for the maintenance of skeletal bone mass during adulthood. Osteolectin binds to Integrin α11 (Itga11), promoting Wnt pathway activation and osteogenic differentiation by leptin receptor+ (LepR+) stromal cells in the bone marrow. Parathyroid hormone (PTH) and sclerostin inhibitor (SOSTi) are bone anabolic agents that are administered to patients with osteoporosis. Here we tested whether osteolectin mediates the effects of PTH or SOSTi on bone formation. We discovered that PTH promoted Osteolectin expression by bone marrow stromal cells within hours of administration and that PTH treatment increased serum osteolectin levels in mice and humans. Osteolectin deficiency in mice attenuated Wnt pathway activation by PTH in bone marrow stromal cells and reduced the osteogenic response to PTH in vitro and in vivo. In contrast, SOSTi did not affect serum osteolectin levels and osteolectin was not required for SOSTi-induced bone formation. Combined administration of osteolectin and PTH, but not osteolectin and SOSTi, additively increased bone volume. PTH thus promotes osteolectin expression and osteolectin mediates part of the effect of PTH on bone formation.

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

12.
Import of yeast peroxisomal matrix proteins is initiated by cytosolic receptors, which specifically recognize and bind the respective cargo proteins. At the peroxisomal membrane, the cargo-loaded receptor interacts with the docking protein Pex14p that is tightly associated with Pex17p. Previous data suggest that this interaction triggers the formation of an import pore for further translocation of the cargo. The mechanistic principles, however, are unclear, mainly because structures of higher-order assemblies are still lacking. Here, using an integrative approach, we provide the structural characterization of the major components of the peroxisomal docking complex Pex14p/Pex17p, in a native bilayer environment, and reveal its subunit organization. Our data show that three copies of Pex14p and a single copy of Pex17p assemble to form a 20-nm rod-like particle. The different subunits are arranged in a parallel manner, showing interactions along their complete sequences and providing receptor binding sites on both membrane sides. The long rod facing the cytosol is mainly formed by the predicted coiled-coil domains of Pex14p and Pex17p, possibly providing the necessary structural support for the formation of the import pore. Further implications of Pex14p/Pex17p for formation of the peroxisomal translocon are discussed.

Peroxisomes are organelles present nearly ubiquitously in eukaryotic cells, ranging from unicellular yeasts to multicellular organisms, such as plants and humans. Beside β-oxidation of fatty acids as a main conserved function of peroxisomes, a broad range of additional metabolic functions is linked to this organelle, underscored by severe and frequently lethal phenotypes of human disorders (1, 2). These organelles do not contain DNA and thus all peroxisomal matrix proteins are encoded in the nucleus and synthesized on free polyribosomes in the cytosol. Subsequently, matrix proteins are targeted to the organelle by peroxisomal import receptors (3). A remarkable feature of peroxisomes is that unlike the transport of unfolded polypeptides across the membranes of the endoplasmic reticulum and mitochondria, they can import already folded, cofactor-bound, and even oligomeric proteins (4, 5). This transport is highly selective and mediated by specific import sequences known as peroxisomal targeting signals (PTSs) (6, 7). Peroxisomal matrix proteins equipped with either a carboxyl-terminal PTS1 or an amino-terminal PTS2, are recognized and bound in the cytosol by the import receptor Pex5p or Pex7p, respectively (8, 9). A peroxisomal membrane-associated complex consisting of Pex13p, Pex14p, and Pex17p in yeast allows docking of the cargo-loaded receptor (1014). This primary interaction of the cargo-loaded receptor with the docking complex induces the formation of a transient and highly dynamic import pore, necessary for the translocation of the cargo across the peroxisomal membrane (1517). How translocation and release of the cargo are realized in detail still remains enigmatic but it has been previously shown that the receptor is exported from the peroxisomal membrane in an ubiquitin- and ATP-dependent manner, a process that is discussed to provide the driving force for cargo import according to the export-driven import model (1820).The receptor–docking complex is of major importance for peroxisomal matrix protein import, as it provides a binding platform for newly formed receptor–cargo complexes at the peroxisomal membrane. Both Pex13p and Pex14p are peroxisomal membrane proteins providing several binding sites for the import receptors Pex5p and Pex7p (16). Docking of the Pex5p–PTS1 protein complex at the peroxisome membrane is supposed to occur at Pex14p (21, 22). Pex17p is tightly associated with Pex14p (23), but its precise function remains unknown. Although Pex17p is part of the docking complex in yeast, it does not significantly contribute to the assembly of the Pex13p/Pex14p subcomplex (15, 23, 24), and its counterpart in higher eukaryotes has not yet been identified. However, Pex17p is essential for peroxisomal import of both PTS1 and PTS2 proteins (14). Strikingly, both import receptors, Pex5p and Pex7p, associate with the docking complex (Pex13p/Pex14p) in absence of Pex17p, but with decreased efficiency (24).Furthermore, albeit a close association between the core components of the docking complex (Pex13p/Pex14p) is important for matrix protein import (25), there are several lines of evidence that Pex13p is not a permanent component of the peroxisomal docking complex or the import pore (10, 26) and interestingly, an assembly between the receptor Pex5p and the docking component Pex14p in absence of Pex13p is capable per se of forming a large transient channel at the peroxisome membrane (15).However, little is known about the molecular mechanism underlying the primary docking and subsequent translocation events, largely because structures of the higher-order assemblies are not available. Here, using cryo-electron microscopy single particle analysis (cryoEM SPA) and cryo-electron tomography (cryoET) combined with cross-linking and native mass spectrometry (MS), we set out to characterize the overall architecture of the yeast Pex14p/Pex17p complex.  相似文献   

13.
Myopia has become a major public health concern, particularly across much of Asia. It has been shown in multiple studies that outdoor activity has a protective effect on myopia. Recent reports have shown that short-wavelength visible violet light is the component of sunlight that appears to play an important role in preventing myopia progression in mice, chicks, and humans. The mechanism underlying this effect has not been understood. Here, we show that violet light prevents lens defocus–induced myopia in mice. This violet light effect was dependent on both time of day and retinal expression of the violet light sensitive atypical opsin, neuropsin (OPN5). These findings identify Opn5-expressing retinal ganglion cells as crucial for emmetropization in mice and suggest a strategy for myopia prevention in humans.

Myopia (nearsightedness) in school-age children is generally axial myopia, which is the consequence of elongation of the eyeball along the visual axis. This shape change results in blurred vision but can also lead to severe complications including cataract, retinal detachment, myopic choroidal neovascularization, glaucoma, and even blindness (13). Despite the current worldwide pandemic of myopia, the mechanism of myopia onset is still not understood (48). One hypothesis that has earned a current consensus is the suggestion that a change in the lighting environment of modern society is the cause of myopia (9, 10). Consistent with this, outdoor activity has a protective effect on myopia development (9, 11, 12), though the main reason for this effect is still under debate (7, 12, 13). One explanation is that bright outdoor light can promote the synthesis and release of dopamine in the eye, a myopia-protective neuromodulator (1416). Another suggestion is that the distinct wavelength composition of sunlight compared with fluorescent or LED (light-emitting diode) artificial lighting may influence myopia progression (9, 10). Animal studies have shown that different wavelengths of light can affect the development of myopia independent of intensity (17, 18). The effects appear to be distinct in different species: for chicks and guinea pigs, blue light showed a protective effect on experimentally induced myopia, while red light had the opposite effect (1822). For tree shrews and rhesus monkeys, red light is protective, and blue light causes dysregulation of eye growth (2325).It has been shown that visible violet light (VL) has a protective effect on myopia development in mice, in chick, and in human (10, 26, 27). According to Commission Internationale de l’Eclairage (International Commission on Illumination), VL has the shortest wavelength of visible light (360 to 400 nm). These wavelengths are abundant in outside sunlight but can only rarely be detected inside buildings. This is because the ultraviolet (UV)-protective coating on windows blocks all light below 400 nm and because almost no VL is emitted by artificial light sources (10). Thus, we hypothesized that the lack of VL in modern society is one reason for the myopia boom (9, 10, 26).In this study, we combine a newly developed lens-induced myopia (LIM) model with genetic manipulations to investigate myopia pathways in mice (28, 29). Our data confirm (10, 26) that visible VL is protective but further show that delivery of VL only in the evening is sufficient for the protective effect. In addition, we show that the protective effect of VL on myopia induction requires OPN5 (neuropsin) within the retina. The absence of retinal Opn5 prevents lens-induced, VL-dependent thickening of the choroid, a response thought to play a key role in adjusting the size of the eyeball in both human and animal myopia models (3033). This report thus identifies a cell type, the Opn5 retinal ganglion cell (RGC), as playing a key role in emmetropization. The requirement for OPN5 also explains why VL has a protective effect on myopia development.  相似文献   

14.
15.
Proper left–right symmetry breaking is essential for animal development, and in many cases, this process is actomyosin-dependent. In Caenorhabditis elegans embryos active torque generation in the actomyosin layer promotes left–right symmetry breaking by driving chiral counterrotating cortical flows. While both Formins and Myosins have been implicated in left–right symmetry breaking and both can rotate actin filaments in vitro, it remains unclear whether active torques in the actomyosin cortex are generated by Formins, Myosins, or both. We combined the strength of C. elegans genetics with quantitative imaging and thin film, chiral active fluid theory to show that, while Non-Muscle Myosin II activity drives cortical actomyosin flows, it is permissive for chiral counterrotation and dispensable for chiral symmetry breaking of cortical flows. Instead, we find that CYK-1/Formin activation in RhoA foci is instructive for chiral counterrotation and promotes in-plane, active torque generation in the actomyosin cortex. Notably, we observe that artificially generated large active RhoA patches undergo rotations with consistent handedness in a CYK-1/Formin–dependent manner. Altogether, we conclude that CYK-1/Formin–dependent active torque generation facilitates chiral symmetry breaking of actomyosin flows and drives organismal left–right symmetry breaking in the nematode worm.

The emergence of left–right asymmetry is essential for normal animal development and, in the majority of animal species, one type of handedness is dominant (1). The actin cytoskeleton plays an instrumental role in establishing the left–right asymmetric body plan of invertebrates like fruit flies (26), nematodes (711), and pond snails (1215). Moreover, an increasing number of studies showed that vertebrate left–right patterning also depends on a functional actomyosin cytoskeleton (13, 1622). Actomyosin-dependent chiral behavior has even been reported in isolated cells (2328) and such cell-intrinsic chirality has been shown to promote left–right asymmetric morphogenesis of tissues (29, 30), organs (21, 31), and entire embryonic body plans (12, 13, 32, 33). Active force generation in the actin cytoskeleton is responsible for shaping cells and tissues during embryo morphogenesis. Torques are rotational forces with a given handedness and it has been proposed that in plane, active torque generation in the actin cytoskeleton drives chiral morphogenesis (7, 8, 34, 35).What could be the molecular origin of these active torques? The actomyosin cytoskeleton consists of actin filaments, actin-binding proteins, and Myosin motors. Actin filaments are polar polymers with a right-handed helical pitch and are therefore chiral themselves (36, 37). Due to the right-handed pitch of filamentous actin, Myosin motors can rotate actin filaments along their long axis while pulling on them (33, 3842). Similarly, when physically constrained, members of the Formin family rotate actin filaments along their long axis while elongating them (43). In both cases the handedness of this rotation is determined by the helical nature of the actin polymer. From this it follows that both Formins and Myosins are a potential source of molecular torque generation that could drive cellular and organismal chirality. Indeed, chiral processes across different length scales, and across species, are dependent on Myosins (19), Formins (1315, 26), or both (7, 8, 21, 44). It is, however, unclear how Formins and Myosins contribute to active torque generation and the emergence chiral processes in developing embryos.In our previous work we showed that the actomyosin cortex of some Caenorhabditis elegans embryonic blastomeres undergoes chiral counterrotations with consistent handedness (7, 35). These chiral actomyosin flows can be recapitulated using active chiral fluid theory that describes the actomyosin layer as a thin-film, active gel that generates active torques (7, 45, 46). Chiral counterrotating cortical flows reorient the cell division axis, which is essential for normal left–right symmetry breaking (7, 47). Moreover, cortical counterrotations with the same handedness have been observed in Xenopus one-cell embryos (32), suggesting that chiral counterrotations are conserved among distant species. Chiral counterrotating actomyosin flow in C. elegans blastomeres is driven by RhoA signaling and is dependent on Non-Muscle Myosin II motor proteins (7). Moreover, the Formin CYK-1 has been implicated in actomyosin flow chirality during early polarization of the zygote as well as during the first cytokinesis (48, 49). Despite having identified a role for Myosins and Formins, the underlying mechanism by which active torques are generated remains elusive.Here we show that the Diaphanous-like Formin, CYK-1/Formin, is a critical determinant for the emergence of actomyosin flow chirality, while Non-Muscle Myosin II (NMY-2) plays a permissive role. Our results show that cortical CYK-1/Formin is recruited by active RhoA signaling foci and promotes active torque generation, which in turn tends to locally rotate the actomyosin cortex clockwise. In the highly connected actomyosin meshwork, a gradient of these active torques drives the emergence of chiral counterrotating cortical flows with uniform handedness, which is essential for proper left–right symmetry breaking. Together, these results provide mechanistic insight into how Formin-dependent torque generation drives cellular and organismal left–right symmetry breaking.  相似文献   

16.
17.
Efficient and faithful replication of the genome is essential to maintain genome stability. Replication is carried out by a multiprotein complex called the replisome, which encounters numerous obstacles to its progression. Failure to bypass these obstacles results in genome instability and may facilitate errors leading to disease. Cells use accessory helicases that help the replisome bypass difficult barriers. All eukaryotes contain the accessory helicase Pif1, which tracks in a 5′–3′ direction on single-stranded DNA and plays a role in genome maintenance processes. Here, we reveal a previously unknown role for Pif1 in replication barrier bypass. We use an in vitro reconstituted Saccharomyces cerevisiae replisome to demonstrate that Pif1 enables the replisome to bypass an inactive (i.e., dead) Cas9 (dCas9) R-loop barrier. Interestingly, dCas9 R-loops targeted to either strand are bypassed with similar efficiency. Furthermore, we employed a single-molecule fluorescence visualization technique to show that Pif1 facilitates this bypass by enabling the simultaneous removal of the dCas9 protein and the R-loop. We propose that Pif1 is a general displacement helicase for replication bypass of both R-loops and protein blocks.

Efficient and faithful replication of the genome is essential to maintain genome stability and is carried out by a multiprotein complex called the replisome (14). There are numerous obstacles to progression of the replisome during the process of chromosome duplication. These obstacles include RNA-DNA hybrids (R-loops), DNA secondary structures, transcribing RNA polymerases, and other tightly bound proteins (59). Failure to bypass these barriers may result in genome instability, which can lead to cellular abnormalities and genetic disease. Cells contain various accessory helicases that help the replisome bypass these difficult barriers (1020). A subset of these helicases act on the opposite strand of the replicative helicase (1, 2, 14, 19).All eukaryotes contain an accessory helicase, Pif1, which tracks in a 5′–3′ direction on single-stranded DNA (ssDNA) (1116). Pif1 is important in pathways such as Okazaki-fragment processing and break-induced repair that require the removal of DNA-binding proteins as well as potential displacement of R-loops (1113, 21, 1518, 2225). Genetic studies and immunoprecipitation pull-down assays indicate that Pif1 interacts with PCNA (the DNA sliding clamp), Pol ε (the leading-strand polymerase), the MCMs (the motor subunits of the replicative helicase CMG), and RPA (the single-stranded DNA-binding protein) (15, 26, 27). Pif1 activity in break-induced repair strongly depends on its interaction with PCNA (26). These interactions with replisomal components suggest that Pif1 could interact with the replisome during replication. In Escherichia coli, the replicative helicase is the DnaB homohexamer that encircles the lagging strand and moves in a 5′–3′ direction (20). E. coli accessory helicases include the monomeric UvrD (helicase II) and Rep, which move in the 3′–5′ direction and operate on the opposite strand from the DnaB hexamer. It is known that these monomeric helicases promote the bypass of barriers during replication such as stalled RNA polymerases (5). The eukaryotic replicative helicase is the 11-subunit CMG (Cdc45, Mcm2–7, GINS) and tracks in the 3′–5′ direction, opposite to the direction of Pif1 (25, 28). Once activated by Mcm10, the MCM motor domains of CMG encircle the leading strand (2932). We hypothesized that, similar to UvrD and Rep in E. coli, Pif1 interacts with the replisome tracking in the opposite direction to enable bypass of replication obstacles.In this report, we use an in vitro reconstituted Saccharomyces cerevisiae replisome to study the role of Pif1 in bypass of a “dead” Cas9 (dCas9), which is a Cas9 protein that is deactivated in DNA cleavage but otherwise fully functional in DNA binding. As with Cas9, dCas9 is a single-turnover enzyme that can be programmed with a guide RNA (gRNA) to target either strand. The dCas9–gRNA complex forms a roadblock consisting of an R-loop and a tightly bound protein (dCas9), a construct that is similar to a stalled RNA polymerase. This roadblock (hereafter dCas9 R-loop) arrests replisomes independent of whether the dCas9 R-loop is targeted to the leading or lagging strand (30). Besides its utility due to its programmable nature (33), the use of the dCas9 R-loop allows us to answer several mechanistic questions. For example, the ability to program the dCas9 R-loop block to any specific sequence enables us to observe whether block removal is different depending on whether the block is on the leading or lagging strand. Furthermore, the inner diameter of CMG can accommodate double-stranded DNA (dsDNA) and possibly an R-loop, but not a dCas9 protein. Using the dCas9 R-loop block allows us to determine the fate of each of its components.Here, we report that Pif1 enables the bypass of the dCas9 R-loop by the replisome. Interestingly, dCas9 R-loops targeted to either the leading or lagging strand are bypassed with similar efficiency. In addition, the PCNA clamp is not required for bypass of the block, indicating that Pif1 does not need to interact with PCNA during bypass of the block. We used a single-molecule fluorescence imaging to show that both the dCas9 and the R-loop are displaced as an intact nucleoprotein complex. We propose that Pif1 is a general displacement helicase for replication bypass of both R-loops and protein blocks.  相似文献   

18.
Genes that are primarily expressed in cochlear glia-like supporting cells (GLSs) have not been clearly associated with progressive deafness. Herein, we present a deafness locus mapped to chromosome 3p25.1 and an auditory neuropathy spectrum disorder (ANSD) gene, TMEM43, mainly expressed in GLSs. We identify p.(Arg372Ter) of TMEM43 by linkage analysis and exome sequencing in two large Asian families segregating ANSD, which is characterized by inability to discriminate speech despite preserved sensitivity to sound. The knock-in mouse with the p.(Arg372Ter) variant recapitulates a progressive hearing loss with histological abnormalities in GLSs. Mechanistically, TMEM43 interacts with the Connexin26 and Connexin30 gap junction channels, disrupting the passive conductance current in GLSs in a dominant-negative fashion when the p.(Arg372Ter) variant is introduced. Based on these mechanistic insights, cochlear implant was performed on three subjects, and speech discrimination was successfully restored. Our study highlights a pathological role of cochlear GLSs by identifying a deafness gene and its causal relationship with ANSD.

Auditory neuropathy spectrum disorder (ANSD) is defined as an inability in speech discrimination despite preserved sensitivity to sound (1). Clinically, ANSD has been characterized by the presence of otoacoustic emission (OAE) and/or cochlear microphonics (CM) and the concurrent absence of averaged auditory brainstem responses (ABR) or presence of abnormal ABR (2, 3). The presence of OAE and/or CM is indicative of normal cochlear outer hair cell (OHC) activity, whereas the abnormal ABR is indicative of disrupted auditory nerve (AN) activity (2). Up to now, research has suggested that ANSD is caused by a malfunctioning of the inner hair cells (IHC), the synapse between the IHCs and AN, or AN itself such as demyelination or desynchronization (46). However, in the organ of Corti of inner ear, glia-like supporting cells (GLSs) comprise major cell types in addition to OHCs and IHCs. GLSs are defined as glia-like cells due to a presence of typical glia markers such as GFAP and GLAST (7). While hair cells play critical roles in mechanoreception and synaptic transmission by converting the acoustic energy into electrochemical signals that are relayed to the brainstem (7), GLSs reside adjacent to hair cells and play critical roles in development and maintenance of auditory system (710). Developing mammalian cochlea can be categorized in two anatomical regions as greater epithelial ridge (GER) and lesser epithelial ridge (LER) (11). The GER refers to the area medial to pillar cells including IHCs and inner GLSs of Kolliker’s organ such as inner phalangeal cells and border cells (12, 13). The LER spans the area radial to pillar cells, consisting of OHCs and outer GLSs such as Hensen’s cells and Deiters’ cells (12). The GLSs in Kolliker’s organ have been studied to generate ATP- and TMEM16A-dependent spontaneous activity that depolarizes IHCs (10, 11, 13), generate spontaneous Ca2+ signaling for OHC refinement (14), and spontaneously regenerate hair cell in the neonatal mouse cochlea (9, 15, 16). Despite their essential function at prehearing stages, the precise role of GLSs in hearing or speech discrimination, especially their contribution to late-onset ANSD, remains unknown.The GLSs are physically coupled to each other by gap junctions (7). Gap junctions consist of two hemi channels, which are encoded by connexin genes, that meet at the plasma membrane of adjacent cells to achieve cell coupling. The connexins provide a pathway for rapid removal of ions from the region of the hair cells during sound transduction (17) by recycling and regulating intracellular K+ and maintaining pH homeostasis (1820). Connexin26 (Cx26, GJB2) and Connexin30 (Cx30, GJB6) are the two predominantly expressed connexins in the GLSs of mammalian cochlea (21) and also the major deafness genes known to induce high incidence of nonsyndromic hearing loss (2124). Variants in connexins break the endocochlear potential of the inner ear and lead to hearing loss (21). Although a pathological role of connexin variants on ANSD has been suggested (25), a combination of heterozygous phenotypes, diverse Cx26 and Cx30 variants (21), and various interacting proteins (26, 27) make it difficult to determine the molecular and cellular mechanism of connexin-related ANSD. Moreover, none of the known interacting proteins for connexins has been associated with ANSD.In this study, we identified two Asian families with hereditary late-onset ANSD with progressive hearing loss. By linkage analysis and exome sequencing, we determined TMEM43-p.(Arg372Ter) variant as the origin of the disease. In order to examine the role of TMEM43, we generated a knock-in (KI) mouse with p.(Arg372Ter) variant which recapitulated a progressive hearing loss with ANSD phenotypes. Ex vivo and in vitro studies further demonstrated that TMEM43 physically interacts with Cx26 and Cx30 in the GLSs. When the p.(Arg372Ter) variant was introduced, passive conductance current in GLSs was disturbed with histological abnormalities in GLSs. Based on the mechanistic insights, cochlear implant (CI) was performed on patients with p.(Arg372Ter) variant, and their speech discrimination was successfully restored. Our study identifies an interacting protein of connexins in cochlea and introduces a role of GLSs in late-onset ANSD.  相似文献   

19.
20.
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