首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 62 毫秒
1.
2.
3.
The role of the kinetochore during meiotic chromosome segregation in C. elegans oocytes has been a matter of controversy. Danlasky et al. (2020. J. Cell. Biol. https://doi.org/10.1083/jcb.202005179) show that kinetochore proteins KNL-1 and KNL-3 are required for early stages of anaphase during female meiosis, suggesting a new kinetochore-based model of chromosome segregation.

Meiosis consists of two consecutive chromosome segregation events preceded by a single round of DNA replication. Homologous chromosomes are separated in meiosis I, which is followed by sister chromatid separation in meiosis II to produce haploid gametes. Both of these stages require chromosomes/chromatids to align during metaphase before separating to opposite poles during anaphase. During mitosis, microtubules emanating from centrosomes at opposite poles of the cell bind chromosomes through a multiprotein complex called the kinetochore, allowing chromosomes to be pulled apart (1, 2). This segregation event takes place in two stages: anaphase A, where chromosomes are pulled toward spindle poles due to microtubule depolymerization, and anaphase B, where spindle poles themselves move farther apart, taking the attached chromosomes with them (3, 4). In many organisms, including mammals, oocytes lack centrosomes, and it has been of great interest to clarify the mechanisms used to ensure chromosomes are properly segregated during female meiosis (5, 6). Caenorhabditis elegans has served as a model for studying both mitosis and meiosis, but the mechanisms operating during female meiosis have been a matter of debate and controversy.In 2010, Dumont et al. showed that the kinetochore is required for chromosome alignment and congression during metaphase (7). However, they suggested that chromosome segregation was the result of microtubule polymerization between the segregating chromosomes (Fig. 1), resulting in a pushing force exerted onto chromosomes toward the spindle poles in a largely kinetochore-independent manner (7). This mechanism was also supported by the finding that CLIP-associated protein (CLASP)–dependent microtubule polymerization between the segregating chromosomes is essential for chromosome separation (8). An alternative model suggested that chromosomes are transported through microtubule-free channels toward the spindle poles by the action of dynein (9). Later evidence put in doubt a role for dynein and favored a model in which chromosomes initially separate when the spindle shortens and the poles overlap with chromosomes in an anaphase A–like mechanism. This is then followed by separation of chromosome-bound poles by outward microtubule sliding in an anaphase B–like fashion (10). However, because microtubules emanating from the spindle poles are not required to separate the homologous chromosomes but microtubules between the separating chromosomes are (8), this model is unlikely, at least as an explanation for mid-/late-anaphase movement. Furthermore, although lateral microtubule interactions with chromosomes predominate during metaphase of C. elegans oocyte meiosis, cryo-electron tomography data described end-on attachments between the separating chromosomes as anaphase progresses (11). This led to the suggestion that lateral microtubule interactions with chromosomes are responsible for the initial separation, but microtubule polymerization between the separating chromosomes is required for the later stages of segregation (11). The mechanisms involved in this initial separation have remained obscure. In this issue, Danlasky et al. show that the kinetochore is in fact required for the initial stages of chromosome segregation during female meiosis—an important step forward in our understanding of the mechanisms governing acentrosomal chromosome segregation (12).Open in a separate windowFigure 1.Some of the key findings in Danlasky et al. Kinetochore proteins surround the outer surface of the chromosomes, resulting in a characteristic cup shape. As anaphase progresses, chromosomes come into close contact to the spindle poles (anaphase A). Chromosome stretching is provided by KNL-1, MIS-12 (KNL-3), and NDC-80 (KMN)–dependent forces. Once the spindle starts elongating (anaphase B), central spindle microtubules provide the pushing forces for chromosome segregation. At this stage, kinetochore proteins also occupy the inward face of separating chromosomes. Upon KMN network depletion, bivalents flatten, and chromosome congression and alignment are defective. Anaphase A chromosome movement is almost absent, which leads to error-prone anaphase B.By simultaneously depleting kinetochore proteins KNL-1 and KNL-3 in C. elegans, Danlasky et al. observed the meiotic chromosome congression and alignment defects described in previous studies (7). However, this double-depletion phenotype displayed three key characteristics that suggested a role for kinetochores in chromosome segregation, which are discussed below.The kinetochore is required for bivalent stretching. It was previously shown that the bivalent chromosomes stretch before the initiation of segregation (10). Danlasky et. al found that this stretching of the chromosomes did not occur when KNL-1,3 were depleted, indicating that the kinetochore is required for this process (Fig. 1). Together with the observation that kinetochore proteins appear to extend toward the spindle poles, this finding suggested that pulling forces resulting from the interaction between the kinetochore and spindle microtubules are occurring during metaphase/preanaphase (Fig. 1).The kinetochore is required for anaphase A. In C. elegans female meiosis, anaphase A occurs when homologous chromosomes begin to separate during spindle shortening, and anaphase B when the chromosomes separate alongside the spindle poles (10). Danlasky et al. observed that KNL-1,3 depletion drastically reduced the velocity of anaphase A, as chromosomes only separated when spindle poles began to move apart. This indicated that pulling forces caused by the interaction between the kinetochore and spindle microtubules are also important for the initial separation of homologous chromosomes in anaphase A.The kinetochore is required for proper separation of homologous chromosomes. In KNL-1,3 depletion strains, 60% of bivalents failed to separate before segregation began, resulting in intact bivalents being pulled to the same spindle pole (Fig. 1). This failure of homologous chromosomes to separate was not thought to be a result of KNL-1,3 depletion interfering with the cleavage of cohesin that holds the two homologous chromosomes together because (a) separase and AIR-2AuroraB, both of which are required for cohesin cleavage, localized normally during metaphase and anaphase, and (b) bivalents separated by metaphase II. This leaves the possibility open that the failure of bivalents to separate was due to the disrupted pulling forces thought to be important in bivalent stretching and anaphase A.Altogether, these data strongly indicate that the kinetochore is required not only for chromosome congression and alignment but also for the early stages of homologue separation. Anaphase B occurred successfully in the absence of KNL-1,3 but was more error prone, likely as a result of the earlier congression and anaphase A defects. While it is clear that chromosome masses do segregate in the absence of the kinetochore, this segregation is highly erroneous as a result of defects during the earlier stages of segregation in anaphase A (Fig. 1).The findings of Danlasky et al. raise testable hypotheses that could significantly enhance our understanding of acentrosomal chromosome segregation. Further investigation of the proposed pulling forces required during metaphase and early anaphase will be of great interest. Additionally, a more detailed analysis of the dynamic localization of separase and Securin, as well as assessing successful cohesin cleavage when KNL-1,3 are depleted, would back up the assertion that the failure of homologous chromosomes to separate was not due to the kinetochore impacting cohesin cleavage. It has previously been shown that the CLASP orthologue CLS-2 in C. elegans localizes to the kinetochore surrounding the bivalent chromosomes during metaphase before relocalizing to the central spindle during anaphase (7, 8, 13). It will be interesting to examine whether this key microtubule-stabilizing protein contributes to anaphase A pulling forces alongside its essential role in microtubule polymerization between chromosomes in anaphase B (8).While the regulation of proper chromosome segregation during acentrosomal meiosis in C. elegans is not yet fully understood, Danlasky et al.’s results represent a significant step forward in this endeavor by showing that the kinetochore is in fact required for the early stages of chromosome segregation.  相似文献   

4.
5.
6.
7.
8.
In this issue, Ayukawa, Iwata, Imai, and colleagues (2021. J. Cell Biol. https://doi.org/10.1083/jcb.202007033) use rapid temporal and high-spatial-resolution electron microscopy imaging to examine the earliest stages of new microtubule nucleation. They discover that straightening of curved tubulin oligomers increases the efficiency of microtubule nucleation.

Microtubules are long cytoskeletal filaments that provide structure to cells, allow for transport within the cells, and participate in cell division by acting together with molecular motors to build a mitotic spindle. While microtubules act as the stiff “bones” of the cell, they also have a unique and important ability to rapidly restructure their length and organization in response to cellular cues. Therefore, large arrays of microtubules, such as in a mitotic spindle, can rapidly depolymerize and disappear as needed. However, in order to rebuild these microtubule networks, the nucleation of new microtubules is required. This nucleation of new microtubules, whether from existing templates such as centrosomes or spindle poles, or via the de novo organization of the tubulin subunits that make up microtubules, remains a poorly understood process.In this issue, Ayukawa, Iwata, Imai, and colleagues used rapid temporal and high-spatial-resolution imaging to study the earliest stages of microtubule nucleation (1). First, the authors purified αβ-tubulin heterodimers with a Y222F mutation in the β-tubulin subunit. This Y222F β-tubulin mutation increased the rate of microtubule assembly and, importantly, greatly accelerated the nucleation rate of new microtubules. Thus, a comparison between wild-type and mutant tubulin allowed the authors to dissect differences in the early nucleation process that could explain the increased nucleation rate for the mutant tubulin. Importantly, a “rapid flush method” was used to capture high-resolution transmission electron microscopy images of tubulin subunits very early in the nucleation process. The rapid flush method revealed “oligomers” of tubulin subunits: chains of tubulin subunits linked together along their long axis.The authors reasoned that these oligomers are likely on-pathway intermediates that are crucial for new microtubule nucleation. Examination of the oligomers revealed differences in curvature between wild-type and mutant tubulin: the mutant tubulin that nucleated more readily had oligomers that were straighter and less curled than the wild-type tubulin. Further, a comparison of GTP tubulin and GDP tubulin early nucleation intermediates revealed that, for both wild-type and mutant tubulin, the GDP tubulin oligomers, which did not nucleate efficiently, were more curved than the GTP tubulin oligomers that nucleated more efficiently (Fig. 1 A, left, red versus blue). Importantly, the fraction of nearly straight oligomers in each sample directly corresponded to the respective nucleation rate for each tubulin type. Thus, the degree of curvature of the oligomers predicted the overall nucleation rate, such that an increase in the fraction of straight oligomers was directly correlated to an increase in the microtubule nucleation rate (Fig. 1 A).Open in a separate windowFigure 1.Long, straight oligomers promote microtubule nucleation. (A) The nucleation of new microtubules is limited by the availability of critical-length, straight GTP tubulin (blue) oligomers. GDP tubulin oligomers (red) are more curved, with reduced nucleation efficiency. (B) Straight GTP tubulin protofilaments (blue) that are attached to templates (gray) could also be required to facilitate the nucleation of new microtubules from templates such as in centrosomes. Curved GDP tubulin protofilaments (red) attached to templates (gray) would not efficiently facilitate nucleation of new microtubules.The authors then examined the role of oligomer length in the nucleation rate, to determine whether there was a critical minimum oligomer length that could predict efficient microtubule nucleation. They fit their bulk microtubule growth curves (turbidity) to a standard nucleation-and-growth model (2) to estimate the minimum size of the oligomers that were likely to grow into microtubules, i.e., the critical length. For wild-type tubulin, ∼4 tubulin dimers were required in order for oligomers to grow into microtubules. However, this critical length was common, and indeed prevalent, within the early nucleation mixtures. Therefore, it seems likely that oligomer curvature, rather than length, is the limiting factor that slows the microtubule nucleation rate (Fig. 1 A).Why would oligomer curvature limit microtubule nucleation rate? While oligomers of sufficient length were readily observed in early nucleation mixtures, lateral association of new oligomers with existing oligomers was rare. Further, when multiple oligomers had indeed associated laterally to form a new, multi-protofilament assembly, the length of the longest strand greatly exceeded the maximum size of the single-stranded oligomers. Thus, it is likely that straight oligomers facilitate the lateral association of new tubulin subunits or oligomers along their length, stabilizing the nascent microtubules and allowing for their stable growth as a multistranded filament.While this work sheds light on the nucleation of new microtubules using purified tubulin, the described results provide interesting insights into potential nucleation mechanisms inside of cells. In cells, microtubules predominately grow from templates, such as centrosomes, and various microtubule-associated proteins may also influence the nucleation process. A recent study of microtubule nucleation from templates found that there was a significant time lag between the arrival of new tubulin subunits to a template and the growth of a new microtubule from the template (3). That study concluded that GTP hydrolysis inhibits microtubule nucleation by destabilizing the nascent microtubule. These results are consistent with the conclusion from Ayukawa, Iwata, Imai, and colleagues that GDP tubulin oligomers are curved and therefore unable to efficiently make lateral associations with other oligomers to stabilize the new microtubule and allow for growth (Fig. 1 B; 1). However, recent electron microscopy studies have also revealed that the GTP tubulin–containing ends of growing microtubules show extended protofilaments with a gentle curvature (4, 5, 6, 7, 8). Thus, one additional barrier to the nucleation of new microtubules on templates may be the straightening of GTP tubulin oligomers. Here, the straightening of template-attached, gently curved GTP tubulin oligomers would then allow for lateral binding of new tubulin dimers and oligomers (Fig. 1 B). The transient straightening of gently curved GTP tubulin oligomers could potentially be accomplished via thermal forces and the resulting protofilament curvature fluctuations. However, this straightening process could also be facilitated by microtubule-associated proteins that promote nucleation, such as TPX2 and XMAP215 (9, 10). In support of this idea, Ayukawa, Iwata, Imai, and colleagues found that the distribution of oligomer curvatures appeared similar to what has been reported for the protofilaments at the growing ends of microtubules (5).In light of these results, interesting future work could explore whether microtubule-associated proteins act to regulate the curvature of the oligomers involved in microtubule nucleation. While many aspects of the mechanisms of microtubule nucleation remain unknown, this new work sheds light on the very earliest stages of microtubule nucleation, which may have wide-ranging implications in future studies of microtubule nucleation and growth in cells.  相似文献   

9.
10.
11.
12.
Spines are tiny nanoscale protrusions from dendrites of neurons. In the cortex and hippocampus, most of the excitatory postsynaptic sites reside in spines. The bulbous spine head is connected to the dendritic shaft by a thin membranous neck. Because the neck is narrow, spine heads are thought to function as biochemically independent signaling compartments. Thus, dynamic changes in the composition, distribution, mobility, conformations, and signaling properties of molecules contained within spines can account for much of the molecular basis of postsynaptic function and regulation. A major factor in controlling these changes is the diffusional properties of proteins within this small compartment. Advances in measurement techniques using fluorescence microscopy now make it possible to measure molecular diffusion within single dendritic spines directly. Here, we review the regulatory mechanisms of diffusion in spines by local intra-spine architecture and discuss their implications for neuronal signaling and synaptic plasticity.

IntroductionNeurons communicate with each other through synapses that organize to create functional circuits. Most excitatory synapses in the central nervous system are formed on dendritic spines, tiny protrusions that extend from dendrites (Bourne and Harris, 2008; Fig. 1 a). The spine typically has a head of 200–1,000-nm diameter, which is connected to the dendritic shaft via a neck of 100–200-nm width (Arellano et al., 2007; Fig. 1 b). The head contains postsynaptic density (PSD) proteins, the actin cytoskeleton, membrane structures, and organelles (Sheng and Hoogenraad, 2007; Fig. 1 c). The molecular composition of spine heads is different from that of the shaft. Because of its characteristic morphology, spines are thought to function as biochemically independent compartments by limiting molecular movement between the spine head and the rest of the dendrite (Adrian et al., 2014; Tønnesen and Nägerl, 2016). Clarifying this regulation is key to understanding how this unitary site of synaptic transmission is controlled. This is particularly crucial to our understanding about how changes in the postsynaptic site lead to synaptic plasticity.Open in a separate windowFigure 1.The shape and internal architecture of dendritic spines. (a) A super-resolution SIM image of a hippocampal neuron dendrite expressing GFP. (b) A surface image of a spine (arrow in a) reconstructed from a SIM image. (c) Schematic representation of a spine containing the PSD, actin cytoskeleton, recycling endosome, and SER.The control of spine architecture is critical at excitatory synapses in the brain (Alvarez and Sabatini, 2007; Forrest et al., 2018). Excitatory synapses exhibit synaptic plasticity, which changes the strength of synaptic transmission through mechanisms at both pre- and postsynaptic sides (Citri and Malenka, 2008). This process is generally thought to be a basis for changes in neural circuits controlled by experiences—i.e., learning and memory (Humeau and Choquet, 2019; Magee and Grienberger, 2020). Here, the size and shape of spines are strongly correlated with the strength of synaptic transmission (Kasai et al., 2010). Spine volume is proportional to PSD area (Harris and Stevens, 1989) and the number of α-amino-3-hydroxyl-5-methyl-4-isoxazole-propionate–type glutamate receptors (AMPARs; Nusser et al., 1998; Matsuzaki et al., 2001). Recently, combinational analysis of electrophysiology and correlative light and EM (CLEM) revealed the linear relationship between PSD area and synaptic strength (Holler et al., 2021). Also, longer spine necks attenuate somatic potentials to a greater degree (Araya et al., 2006). Thus, structural and functional plasticity of spines is tightly regulated. Specifically, when synaptic transmission is strengthened (e.g., long-term potentiation [LTP]), spines grow (Matsuzaki et al., 2004). In turn, when synaptic transmission weakens (e.g., long-term depression), spines shrink (Zhou et al., 2004; Oh et al., 2013).While many molecules involved in the plasticity of spine synapses have been identified (Sala and Segal, 2014), their mechanisms and regulations can only be discovered by monitoring the regulated changes in the composition and signaling properties of these factors within the confined space of the spine’s cytoplasm. To this end, the development of local photolysis of caged-glutamate played an important role (Matsuzaki et al., 2001). This method made it possible to induce structural plasticity locally at a single spine. Spine enlargement is induced by uncaging of caged-glutamate in the absence of Mg2+ or with postsynaptic depolarization in the presence of Mg2+ to activate N-methyl-D-aspartate–type glutamate receptors (NMDARs; Matsuzaki et al., 2004). Conversely, spine shrinkage is induced by low-frequency uncaging of caged-glutamate in the absence of Mg2+ or with postsynaptic depolarization (Oh et al., 2013). Shrinkage can also be induced by glutamate uncaging temporally coupled with back propagation action potential and uncaging of caged–γ-aminobutyric acid (GABA; Hayama et al., 2013). Glutamate uncaging–induced structural plasticity has also been seen to occur in vivo (Noguchi et al., 2019).These methods have vastly improved our understanding of the molecular mechanisms of structural plasticity (Nishiyama and Yasuda, 2015). In particular, stimulus-dependent increases in spine size (structural LTP [sLTP]), which are thought to be associated with functional LTP, have been studied extensively as a model of LTP (Nakahata and Yasuda, 2018; Fig. 2). Strong synaptic input causes an influx of Ca2+ through NMDARs that activates Ca2+/calmodulin-dependent protein kinase II (CaMKII) and the downstream signaling cascades. This modification of signaling cascades can affect cytoskeletal organization and membrane trafficking, which are responsible for two subsequent cellular events. First, spine morphology is modulated through cytoskeletal changes (Borovac et al., 2018). Second, synaptic transmission is enhanced by increased AMPAR insertion into the plasma membrane and movement to the PSD (Huganir and Nicoll, 2013). However, the molecular mechanisms that link these two phenomena are not fully understood (Herring and Nicoll, 2016). As sLTP progresses, the molecular composition within spines changes (Bosch et al., 2014; Meyer et al., 2014). Specifically, immediately after sLTP induction, actin-related molecules such as cofilin and actin-related protein 2/3 (Arp2/3) complex accumulate within a stimulated spine. On the other hand, scaffold proteins such as PSD-95 slowly accumulate over tens of minutes. SynGAP, which is localized at the PSD through interaction with PSD-95, escapes from spines immediately after stimulation and contributes to the expression of sLTP (Araki et al., 2015). These changes in molecular compositions immediately after stimulation may be due not only to molecule-specific binding but also to the physical regulation of diffusion (Obashi et al., 2019).Open in a separate windowFigure 2.Molecular motion important in sLTP. Strong synaptic input causes an influx of Ca2+ through NMDARs that activates CaMKII and the downstream signaling cascades. This modification of signaling cascades can affect cytoskeletal organization and membrane trafficking, which regulate spine morphology. Spine morphology affects the molecular exchange between the spine head and the dendritic shaft and lateral diffusion of membrane proteins including AMPARs. Regulation of molecular movements through the spine neck affects the molecular composition within spines. This change affects signal propagation into nearby spines. For example, cofilin and Arp2/3 complex accumulate within spines. SynGAP and activated RhoA escape from spines. Reorganization of the actin cytoskeleton affects movement of large molecules and the formation of a large signaling complex containing CaMKII and Tiam1. Also, the structure of the PSD affects membrane protein diffusion and alters the synaptic trafficking of AMPARs.In addition to molecular localization, fluorescence lifetime imaging of FRET-based biosensors has made it possible to measure spatiotemporal changes in the activity of signaling molecules involved in sLTP (Yasuda, 2012). These studies have demonstrated a critical relationship between the time that a molecule spends within a spine and the rate of signal inactivation. This relationship determines whether an activated signaling molecule is confined within a single spine or escapes from the spine and interacts with effectors present in the adjacent dendritic shaft or nearby spines (Yasuda, 2017). The signal propagation into nearby spines is most likely related to heterosynaptic plasticity, where activated synapses influence neighbor synapses within the same dendritic segments (Oh et al., 2015; Colgan et al., 2018; Chater and Goda, 2021). Thus, diffusion is a central feature of the regulation of spine structural plasticity. However, because the size of spines is small relative to the spatial resolution of diffraction-limited fluorescence microscopy and measuring methods are limited, elucidation of the mechanism regulating diffusion within spines has been challenging.To address this gap in understanding, researchers advanced fluorescence microscopy techniques, which enabled us to measure changes in the nanoscale localization, diffusion, and signal activities inside spines. These studies allow us to directly understand how spine structures physically limit molecular diffusion and reveal fundamental mechanisms that control the localization and biochemical signal transduction pathways in neurons. Here, we summarize recent findings that have revealed physical barriers within spines using super-resolution microscopy (Sigal et al., 2018; Table 1) and molecular dynamics measurements (Fig. 3 and Table 2), and we discuss how these barriers serve as a fundamental feature controlling neuronal signaling and synaptic plasticity.Table 1.List of super-resolution microscopy techniques
TechniquePrincipleResolutionCommentsApplications in spines and synapses
Lateral (XY)Axial (Z)
CLSM250 nm500 nm
TPLSMTwo-photon excitation350 nm700 nmDeeper tissue penetration; adaptive optics further improveHelmchen and Denk, 2005; Ji, 2017
STEDStimulated emission (Vicidomini et al., 2018)20–70 nm500 nm3-D STED increases axial resolution; chronic in vivo imaging is possibleNägerl et al., 2008; Berning et al., 2012; Pfeiffer et al., 2018
SIMMoiré effect with structured illumination (Wu and Shroff, 2018)100 nm250 nmNo need for special fluorophores; limited resolution improvementKashiwagi et al., 2019; Li et al., 2020
SMLM (PALM, STORM)Photoactivation, photoconversion (Baddeley and Bewersdorf, 2018)10–30 nm30–60 nmHigh spatial resolution; temporal resolution is relatively worseDani et al., 2010; Tang et al., 2016
ExMPhysical expansion of sample (Wassie et al., 2019)4–20-fold improvement4–20-fold improvementCapable of combining with other imaging techniques, only for fixed samplesGao et al., 2019; Sarkar et al., 2020 Preprint
Open in a separate windowCLSM, confocal laser scanning microscopy; SMLM, single-molecule localization microscopy; STORM, stochastic optical reconstruction microscopy; TPLSM, two-photon laser scanning microscopy.Open in a separate windowFigure 3.Imaging techniques to measure diffusion inside dendritic spines. (a) FRAP. Fluorescence intensity change is measured after photobleaching fluorescent molecules in a spine head. Fluorescence recovery rate is mostly determined by the exchange rate between spine and dendrite. (b) FCS. The fluctuation of fluorescence intensity from the detection volume fixed inside a spine head (blue region in left panel) is recorded as a function of time (center panel). Since the fluorescence intensity fluctuates as the molecules enter and leave the fixed detection volume, the characteristics of intensity fluctuation essentially contain information about local diffusion speed. To estimate the diffusion coefficient, the autocorrelation function of fluorescence intensity fluctuation is calculated (right panel). (c) SPT. In SPT, molecular trajectory is directly measured with video microscopy. To analyze the speed and pattern of molecular motion, mean squared displacement (MSD) is calculated. For diffusion without barrier, MSD increases linearly against time. On the other hand, for diffusion within the compartment, MSD converges to a certain value, which corresponds to compartment size. (d) Comparison of three measurement techniques.Table 2.List of fluorescence molecular dynamic measurement techniques
TechniquePrincipleApplications in spines and synapses
FRAPFluorescent molecules in a small region are photobleached, and subsequent movement of surrounding nonbleached fluorescent molecules into the photobleached area is monitored (Lippincott-Schwartz et al., 2018).Svoboda et al., 1996; Bloodgood and Sabatini, 2005
FCSFluctuation of fluorescence intensity from the detection volume fixed at a specific position is recorded, and a temporal correlation is analyzed (Elson, 2011).Chen et al., 2015; Obashi et al., 2019
RICSSpatial correlation is analyzed from raster-scanned images (Digman and Gratton, 2011).Obashi et al., 2019
SPTThe movement of a single particle is tracked using time-lapse imaging, and a trajectory is made and analyzed. To detect single particles, the density of fluorescence particles should be kept low (Shen et al., 2017).Borgdorff and Choquet, 2002; Varela et al., 2016
SPT-PALMOnly a small number of photoactivatable fluorescent proteins in the field of view are activated and tracked until they are bleached (Manley et al., 2010).Frost et al., 2010b; Nair et al., 2013
Open in a separate windowSpine structures and diffusionThe complex physical structures of spines can impact the diffusion of molecules inside the spine cytoplasm and between spines and their parental dendritic shafts (Fig. 1 c and Fig. 2). For example, consider the diffusional translocation of molecules between the PSD and dendritic shaft. For cytoplasmic proteins, because a spine is connected to the dendritic shaft through a narrow neck, proteins must pass through the neck by diffusion or slow active transport. A spine neck functions as a diffusion barrier because of its narrow width (Svoboda et al., 1996). Molecular complexes with actin filaments and related proteins, such as synaptopodin and ankyrin-G, that maintain this characteristic neck morphology may also affect diffusion. Furthermore, the cytoplasm within spines is likely to be not homogeneous but organized with multiple nanoscale domains with different biophysical properties (Frost et al., 2010a; MacGillavry and Hoogenraad, 2015). Thus, these locally dense cytoskeletal and membranous structures can limit the molecular path of diffusion within a spine by specific binding interactions or nonspecific local steric effects. These factors will change the residence time of proteins within spines.Besides cytosolic proteins, membrane protein diffusion can be regulated by structures on and near the plasma membrane. For example, the cortical cytoskeleton affects the movement of membrane proteins (Kusumi et al., 2012). Furthermore, specialized membrane domains with a high density of membrane-associated structures, such as synaptic contact sites, accumulate many relatively immobile molecules and limit membrane protein diffusion (Trimble and Grinstein, 2015). Lastly, spines are not simply spherical. Boundaries between the spine shaft and neck—and also the spine head and neck—can contain high curvatures. Also, large spine heads contain a concave surface (Kashiwagi et al., 2019). Thus, local concavities, undulations, and convexities may affect the possible path a molecule can take (Simon et al., 2014; Klaus et al., 2016). From all these factors, the shape and internal architecture of spines can have strong effects on diffusion for both cytosolic and membranous proteins.Influences of spine morphology on diffusional coupling between spines and dendritesAlthough the cytoplasm of spines is directly connected to the cytoplasm of dendritic shafts, a narrow neck is thought to limit diffusion of both cytosolic and membrane molecules between two compartments (Holcman and Schuss, 2011; Kusters et al., 2013; Ramirez et al., 2015). FRAP is a method that can be used to measure the diffusional speeds from an exchange rate between nonbleached and bleached molecules after bleaching fluorescent molecules in a small region (Lippincott-Schwartz et al., 2018; Fig. 3, a and d). Local photoactivation or photoconversion and subsequent measurements of fluorescence intensity is another technique comparable to FRAP (Bancaud et al., 2010). When fluorescence bleaching is performed in a spine head, the speed of fluorescence recovery mostly reflects the rate of molecular exchange between the head and the connected dendrite. Because the diffusion of small molecules in a head is faster than the rate of molecular exchange between spines and dendrites, the fast component of intra-spine diffusion is more difficult to detect in FRAP recovery curves (Svoboda et al., 1996). FRAP or photoactivation experiments of cytoplasmic and membrane-anchored fluorescent proteins showed that diffusional coupling between spines and dendrites varies between spines (Bloodgood and Sabatini, 2005; Ashby et al., 2006). Since the shape of spines is diverse, it has been proposed that this diversity underlies variability in spine–dendrite coupling. However, because the details of spine morphology cannot be analyzed with the spatial resolution of diffraction-limited fluorescence microscopy, a relationship between the shape of spines and diffusional coupling had not been directly demonstrated.Recently, however, super-resolution microscopy has made it possible to analyze spine shape in living neurons with a spatial resolution of ∼50 nm (Nägerl et al., 2008). Influences of spine morphology on diffusional coupling were verified experimentally for the first time by directly comparing the morphological features of spines and diffusional coupling. This comparison was achieved by stimulated emission depletion (STED) microscopy of spines combined with FRAP of YFP or Alexa dyes applied to the same spine (Takasaki and Sabatini, 2014; Tønnesen et al., 2014). These direct comparisons indicated that the diversity in diffusional couplings could be explained solely by the diversity of spine shapes for more than half of the measured spines. In other words, it was shown that for many spines, the exchange rate (τ) of small molecules within spines could be explained by a single-compartment model (Svoboda et al., 1996) described by the shape of the spine:τ=V×LA×D ,where V is the head volume, L is the length of the neck, A is the cross-sectional area of the neck, and D is the diffusion coefficient of molecules. Also, sLTP induction made a spine neck thicker and shorter (Tønnesen et al., 2014). This change in the spine neck complements the decrease in the coupling rate associated with the increase in the spine head volume. This coordinated morphological change appears to maintain molecular concentration in a spine.Besides the work focusing on cytoplasmic proteins, the influence of spine shape on the diffusional coupling of membrane molecules has also been investigated (Adrian et al., 2017). The spine–dendrite diffusional coupling was tested by photoactivated localization microscopy (PALM) and photoconversion experiments using membrane-anchored mEos3.2 as a probe. This study showed that even if spines have the same surface area and neck width, the diffusional coupling varies between different spine shapes. Therefore, a model spine was created based on the experimentally measured spine shape parameters, and a simulation was conducted on the model spine and compared with the experiment. As a result, although experimental results tended to provide slower diffusion kinetics than simulation values, experiments showed a good correlation with simulations based on the spine shape parameters alone.Experiments have confirmed that spine morphology is a major factor determining the diffusional coupling for both cytoplasmic and membrane-bound molecules in dendrites. However, for some spines, the simulated and experimental results diverge. One possibility is that the effects of local intra-spine architectures on molecular diffusion vary for each spine. Another possibility is that there was insufficient spatial resolution for reconstructing the spine morphology. Although the above studies used rotationally symmetric shapes as model spines, actual spines are not rotationally symmetrical structures and generally have a more complicated morphology and surface features (Nägerl et al., 2008; Berning et al., 2012; Kashiwagi et al., 2019; Zaccard et al., 2020; Fig. 1, a and b). Thus, it is possible that estimations of spine shape were insufficient or that the fine structure of spines affects diffusion. In this regard, developing an analysis method for spine morphology from both the experimental and computational sides is key (Okabe, 2020a; Tamada et al., 2020). Recently, Kashiwagi et al. (2019) developed a 3-D structured illumination microscopy (SIM)–based nanoscale analysis of spine morphology. Direct comparison of SIM images and serial-section EM images revealed that the basic morphological features were highly correlated among these images. This indicates the high precision of SIM-based nanoscale spine analysis. To analyze spines computationally, SIM images were converted into a computational geometry, and morphological features were calculated. Then, these features were analyzed by principal component analysis. By mapping the temporal changes of spine morphology obtained by live-cell SIM imaging in the dimension-reduced feature space, the authors revealed that the spine population can be categorized based on different simplified morphological dynamics.Also, expansion microscopy (ExM) is another new and important imaging technique for spine structural analysis (Wassie et al., 2019). However, it can only be applied to fixed samples. Since ExM samples are transparent, 3-D super-resolution imaging is available for thick samples with large volumes (Gao et al., 2019). With recent developments in sample preparation technology, ExM has the potential to investigate spine morphology and localization of multiple biomolecules and organelles within a single sample (Chozinski et al., 2016; Tillberg et al., 2016; Karagiannis et al., 2019 Preprint; Sun et al., 2021). Minimizing the distortion of isotropy during expansion will be important for nanoscale morphological analysis. In the future, combining dynamic fluorescence measurements and structural measurements gained from EM (CLEM) will be a powerful approach to evaluate the effects of spine ultrastructure on molecular diffusion in greater nanoscale detail (Maco et al., 2013; Taraska, 2015; Luckner et al., 2018).Along with biochemical compartmentalization, dendritic spines have been proposed to be important for electrical compartmentalization (Yuste, 2013; Araya, 2014; Tønnesen and Nägerl, 2016). Spine morphology, particularly spine neck morphology, is thought to be critical for this effect (Cartailler et al., 2018). Several studies have sought to measure neck resistance based on morphological analysis using EM (Harris and Stevens, 1989; Tamada et al., 2020), super-resolution microscopy (Tønnesen et al., 2014), FRAP of small molecules (Svoboda et al., 1996; Tønnesen et al., 2014), glutamate uncaging (Araya et al., 2006; Takasaki and Sabatini, 2014), calcium imaging (Grunditz et al., 2008; Harnett et al., 2012), voltage imaging (Popovic et al., 2015; Acker et al., 2016; Kwon et al., 2017), and intracellular recordings directly from spine heads (Jayant et al., 2017). However, results were not completely consistent, and the degree of electrical compartmentalization is still unclear. Thus, the relationship between spine morphology and electrical signaling of the synapse is still an open question. Likewise, how morphological changes in the neck induced by LTP affect dendritic computation will be an important area of future study (Araya et al., 2014; Tazerart et al., 2020).Actin cytoskeletonThe cytoskeleton in spines is primarily composed of actin (Hotulainen and Hoogenraad, 2010; Okabe, 2020b). Actin is present in high densities in both the head and neck regions (Korobova and Svitkina, 2010). Actin polymers are essential in controlling the localization of PSD molecules and in changing and maintaining spine morphology (Frost et al., 2010a; Bertling and Hotulainen, 2017). In addition to these functions, dense actin polymers in spines may regulate synaptic functions by controlling diffusion because the intracellular cytoskeleton and membrane structures influence diffusion (Novak et al., 2009). If this regulation occurs in spines, variations in the distribution of intra-spine structures can be a factor in the large deviations between the measured values of diffusional coupling and the value predicted from models. A ratio of the spine FRAP recovery time of Alexa dyes to that of YFP was comparable to that of hydrodynamic radii (Tønnesen et al., 2014). This suggests that the suppression of diffusion by actin polymers is weak for molecules with the size of GFP. However, suppressive effects on molecular diffusion by the cytoskeleton, such as actin polymers, is dependent on the size of molecules (Baum et al., 2014; Katrukha et al., 2017). Thus, diffusion of larger molecules may be influenced to a greater degree by actin polymers.Because the shape of spines affects the recovery time of FRAP measurement, it is difficult to investigate the effects of intra-spine structure on molecular diffusion using FRAP alone. Therefore, there is a need for a method capable of measuring diffusion directly in confined spaces. Lu et al. (2014) measured the motion of mEOS2-fused CaMKIIα in spines by single-particle tracking (SPT)–PALM. SPT can directly evaluate diffusional speed in spines because it analyzes the molecular movement trajectory of single molecules (Fig. 3, c and d). The SPT measurement showed that CaMKIIα exhibited at least three different diffusion modes within spines: (1) a free diffusion component, (2) a component bound to immobile molecules, and (3) a component moving at an intermediate velocity. Depolymerization of actin polymers by latrunculin A reduced the proportion of molecules with intermediate velocities in spines while concomitantly increasing the free diffusion component. Also, diffusional speeds of CaMKIIα were slower and the ratio of the intermediate component was larger in spines than in dendrites. Because the transition between free and bound states would occur rarely during the measurement period due to the slow unbinding rate of CaMKII from actin polymers, transient binding alone does not explain the mechanism for the intermediate velocity. Although the details are unclear, CaMKII motion is restricted by actin polymers through a mechanism distinct from direct binding, including a molecular sieve effect or transient binding to actin-associated molecules.Obashi et al. (2019) used fluorescence correlation spectroscopy (FCS) and raster image correlation spectroscopy (RICS) to measure the diffusion of biologically inert probes within spines. FCS is a method for estimating diffusion speed from the time taken for fluorescent molecules to pass through the detection volume excited by a high numerical aperture objective and is capable of measuring fast diffusion within a small cellular compartment (Elson, 2011; Fig. 3, b and d). RICS is another method for estimating diffusion speed from the spatial similarity of fluorescence intensity in a scanned image (Digman and Gratton, 2011). Since FCS and RICS are affected by the small size of spines due to the boundary effect, it is not possible to measure the diffusion coefficient accurately (Jiang et al., 2020). Still, by averaging, values proportional to the actual values can be obtained. Diffusion of GFP and GFP tandem pentamer (GFP5) were compared, and only diffusion of GFP5 within spines was enhanced by depolymerizing actin with latrunculin A treatment. Molecular dynamics simulation confirmed that the diffusion of molecules over the size of GFP5 was suppressed by actin polymers with a density (380 µM) estimated from the values in the literature and experiments.Together, these experiments support the idea that a meshwork of dense actin polymers in spines acts as a physical barrier to the diffusion of larger (>100 kD) molecules (Fig. 4 a). Photoactivation experiments of intra-spine photoactivatable GFP (PA-GFP)–actin showed that there are at least three groups of actin polymers with different reorganization rates (Honkura et al., 2008). In addition, experiments with SPT-PALM of PA-GFP–actin showed that a rate of actin filament polymerization increased near PSDs (Frost et al., 2010b). PALM analysis also revealed that actin-related molecules within spines are arranged in a manner specific for each molecule (Chazeau et al., 2014). These results suggest that the diffusional control by actin polymers in spines may differ between each subcompartment.Open in a separate windowFigure 4.Diffusion within network of actin polymers and PSD. (a) Comparison of the size of actin polymer network and diffusion molecules. Average distance between actin polymers is estimated for actin polymers with 380 µM (Obashi et al., 2019). GFP is represented as a diameter of 3 nm and CaMKII is represented as a diameter of 20 nm (Myers et al., 2017). (b) A schematic model of AMPAR diffusion within a crowded PSD. Different localization patterns of molecules cause different diffusion patterns. Such mechanisms will occur within the PSD. Density and size of molecules are based on the literature (Okabe, 2007; Li et al., 2016).It was also shown that reorganization of the actin cytoskeleton immediately after sLTP induction (Chazeau and Giannone, 2016; Mikhaylova et al., 2018) enhanced the diffusion of larger molecules within the spine head (Obashi et al., 2019). Further, FRAP experiments showed that diffusional coupling and synaptic translocation of large synaptic molecules, such as CaMKII and T cell lymphoma invasion and metastasis-inducing protein 1 (Tiam1), were facilitated at the initial phase of sLTP. Thus, the reorganization of actin polymers regulates molecular translocation between dendrites and the PSD in coordination with morphological changes of the spine neck (Tønnesen et al., 2014). The enhancement of molecular diffusion by actin may also be related to the formation of a large signaling complex containing CaMKII and Tiam1 and may be an important physical mechanism responsible for the initiation of sLTP (Saneyoshi et al., 2019).Membranous organellesAlong with the cytoskeleton, dendrites also contain many membranous organelles and compartments, and some are present in spines (Bourne and Harris, 2008). Smooth ER (SER) and recycling endosomes are present in <50% of spines. Spine apparatus, which is composed of stacked SER, is present in 10–20% of spines. Localizations of these organelles change after LTP induction, and spines containing SER are larger than those without SER (Chirillo et al., 2019; Kulik et al., 2019; Perez-Alvarez et al., 2020). While mitochondria are abundant in dendritic shafts but rarely present in spines (Wu et al., 2017), synaptic activation relocates mitochondria into spines (Li et al., 2004). Therefore, variations in the diffusional coupling between spines and dendrites could be due to the heterogeneous localization of these organelles (Cugno et al., 2019). Holbro et al. (2009) compared the diffusional coupling of ER-containing spines and ER-free spines by using an ER-targeted GFP probe. FRAP recovery times of RFP were not different among ER-containing and ER-free spines, indicating that the ER does not block cytoplasmic diffusion between spines and dendritic shafts. Understanding both spine morphology and the volume of ER within spines in the future will clarify the effects of excluded volume by the ER and other organelles in more detail.Structures around spine necksMolecules present in spine necks may physically control diffusion by forming a complex higher-order structure. Platinum replica EM showed the presence of Arp2/3 complex within necks and a longitudinal network of branched and linear actin filaments (Korobova and Svitkina, 2010). SPT-PALM of PA-GFP–actin also showed that actin polymers in necks are dynamically reorganized and that they are arranged in many orientations (Frost et al., 2010b). Thus, actin polymers in spine necks may affect molecular diffusion. Synaptopodin, for example, is an actin-binding protein located predominantly in spine necks. It is colocalized with the spine apparatus (Vlachos, 2012). Wang et al. (2016) used SPT to measure metabotropic glutamate receptor 5 (mGluR5) diffusion around necks. They compared the diffusion of mGluR5 around the necks of spines containing (or not containing) synaptopodin. The diffusion of mGluR5 decreases around spine necks near synaptopodin clusters. Further, latrunculin A treatment specifically enhanced the diffusion around spine necks near synaptopodin clusters. These results suggest that synaptopodin regulates the actin polymer network around spine necks. This actin complex can act as a diffusion barrier for membrane proteins.Another protein that has been implicated in diffusional control of membrane proteins is ankyrin-G. Ankyrin-G forms nanodomain at perisynaptic membranes and in spine necks (Smith et al., 2014). AMPARs accumulated in spines with ankyrin-G clusters and showed slower spine–dendrite coupling. Ankyrin-G is the major cytoskeletal scaffold of the axon initial segment (AIS; Leterrier, 2018). Ankyrin-G and actin scaffolds densely accumulate at the AIS and inhibit diffusion of the membrane and cytoplasmic molecules (Winckler et al., 1999; Nakada et al., 2003; Song et al., 2009). Also, super-resolution microscopy recently revealed the presence of membrane-associated periodic skeleton composed of actin rings, spectrin, and accompanying proteins in the axon including the AIS (Xu et al., 2013; Zhong et al., 2014). At the AIS, the actin rings and associated structures act as a diffusion barrier to membrane proteins (Albrecht et al., 2016). Adding to the axon, membrane-associated periodic skeleton was also observed in the dendrites and spine necks (Bär et al., 2016; Sidenstein et al., 2016). Therefore, it is interesting to postulate that a molecular complex with actin filaments similar to the AIS is also present in spine necks and could regulate molecular diffusion in this small compartment.Another cytoskeletal component, septin 7, localizes to the base of spines and acts as a diffusion barrier for membrane-bound molecules (Ewers et al., 2014). Recently, actin patches were found at the base of spines and were shown to be remodeled by synaptic activity. These structures modulate microtubule entry into spines and the transport of lysosomes (Schätzle et al., 2018; van Bommel et al., 2019). It is interesting to ask whether actin patches at spine bases affect molecular diffusion. There are still many unknown features at the spine neck, and how these structures limit the diffusion of cytoplasmic and membrane molecules to control neuronal functions remains unclarified.Molecular crowding in the PSDThe PSD is a membrane-associated structure containing densely packed postsynaptic molecules (Sheng and Hoogenraad, 2007). It was originally identified as an electron-dense structure in EM (Okabe, 2007). The number and location of receptors and adhesion molecules in PSDs are directly related to synaptic function (Chen et al., 2018). SPT studies indicate that AMPARs diffuse laterally into and out of PSDs and regulate synaptic function by controlling the number and location of AMPARs (Choquet and Hosy, 2020). Because there are many scaffold proteins in PSDs, membrane proteins including AMPARs accumulate in PSDs due to intermolecular binding. Furthermore, because the molecular density in PSDs is high, the accumulation of membrane proteins may be regulated by the suppression of mobility within the PSD and molecular exchange at the boundary of PSDs (Gerrow and Triller, 2010; Kokolaki et al., 2020).To check this possibility, Li et al. (2016) combined FRAP, SPT, and Monte Carlo simulation to investigate the effect of molecular crowding of PSDs on the lateral diffusion of membrane molecules. When the intracellular domain size of membrane proteins was large, diffusion within the PSD and the exchange rate between the inside and outside of the PSD decreased. Super-resolution microscopy showed that the distribution of PSD-95, a major scaffolding protein of the PSD, within PSDs is not uniform (Fukata et al., 2013; MacGillavry et al., 2013; Nair et al., 2013; Broadhead et al., 2016; Gwosch et al., 2020). Interestingly, the simulation showed that the residence time of membrane proteins within PSDs was longer in the condition of experimentally measured PSD-95 distribution, while the residence time decreased with a random distribution of PSD-95 (Li et al., 2016).Recently, the shape of PSDs inside spines induced by sLTP was analyzed by CLEM (Sun et al., 2019 Preprint). It was shown that rearrangements of PSD shape occurred immediately after induction of sLTP (<3 min), and the PSD took more complex morphology. This increased structural complexity persisted in the late phase (120 min). PSD size and the accumulation of PSD-95 increased slowly over several tens of minutes after sLTP induction (Meyer et al., 2014), whereas synaptic transmission efficiency increased immediately (Matsuzaki et al., 2004). This difference in time may be explained by a mechanism in which the acute ultrastructural changes of the PSD without net growth of the molecular assembly alter the mobility of AMPARs by changing the distribution of a physical barrier, leading to alternations in the number and localization of AMPARs (Fig. 4 b). In future studies, it will be necessary to clarify how coordination between intermolecular binding and physical diffusion barriers in PSDs supports both acute accumulation of AMPARs and their subsequent stabilization in stimulated spines. Further, 3-D SIM imaging revealed that the concave surface of the spine head, which interacts with presynaptic membranes, is enlarged and stabilized by sLTP induction (Kashiwagi et al., 2019). In the future, it will be interesting to determine the relationships between concave membrane surfaces, PSD morphologies, and the dynamics of receptors and adhesion molecules at single spines.In addition, although AMPAR has been thought to be present as a tetramer (Greger et al., 2007), recent observations of SPT have shown that the majority of diffusive AMPARs are monomers or dimers (Morise et al., 2019). Molecular diffusion in the monomer form increases an exchange rate between the inside and the outside of PSDs, making it possible to efficiently change the AMPAR composition within synapses. It remains to be seen whether other molecular complexes, such as NMDARs and cell adhesion molecules, also modulate their diffusion within the molecularly dense PSD by changing their oligomeric state.Conclusion and outlookHere, we have highlighted key recent findings on the relationship between molecular diffusion and physical barriers within spines. The regulation of molecular diffusion is important for sLTP expression (Fig. 2). Spine structural changes during sLTP will affect synaptic function in a coordinated manner (Fig. 5). For example, after sLTP induction, the actin network is reorganized and diffusion of large molecules is enhanced (Obashi et al., 2019). This facilitates the formation of large signaling complexes and the rearrangement of protein complexes within spines. At the same time, spine necks become wider and shorter, and spine heads enlarge (Tønnesen et al., 2014). Changes in actin and spine morphology enhance the molecular movement between the PSD and the shaft and are important for the relocation of proteins (Fig. 2). These structural changes occur in the early phase of sLTP. Thus, the cooperative regulation of diffusion might act as a precise temporal switch of sLTP induction. Also, this enhancement of molecular exchange affects the relocation of activated signaling molecules into the shaft or nearby synapses, which leads to heterosynaptic plasticity (Yasuda, 2017). Potentiation of synaptic transmission requires synaptic trafficking of AMPARs (Choquet and Hosy, 2020). Although both spine morphology (Adrian et al., 2017) and PSD structure (Li et al., 2016) affect membrane protein diffusion, how structural changes associated with sLTP induction affect diffusion will be clarified in the future. Furthermore, the effects of transient SER visits (Perez-Alvarez et al., 2020) and structural changes around spine necks are an important area for future work. Although the relationship between structure and diffusion in sLTP is critical, the difficulty of measurements with a small single spine has made a comprehensive view difficult to obtain. Thus, future work will be necessary to clarify how structural changes affect diffusion and how this physical change to dendritic spines cooperatively modulates synaptic functions.Open in a separate windowFigure 5.Changes in the shape and internal architecture of spines after induction of sLTP. At the initial phase of sLTP, a spine head expands. In addition, the spine neck becomes wider and shorter (Tønnesen et al., 2014), and a concave surface area of spine head is increased (Kashiwagi et al., 2019). The actin polymer network is reorganized (Obashi et al., 2019), and the SER visits within a spine transiently (Perez-Alvarez et al., 2020). Also, PSD shape becomes more complex (Sun et al., 2019 Preprint). These physical changes should occur in concert and will affect molecular composition and biochemical signaling through diffusional regulation. These physical changes will act as a precise temporal switch of sLTP induction.Although new imaging techniques have demonstrated the connection between diffusion and physical barriers, little is known about how changes in the movement of molecules alter synaptic functions (Reshetniak et al., 2020b). Because of the small volume of the spine, very small molecules with high diffusivity, such as Ca2+, are expected to spread rapidly (∼1 ms) by diffusion (Chen and Sabatini, 2012). For large molecules such as signaling complexes, it remains to be seen whether spatially uniform diffusion takes place or whether local heterogeneity in the spine cytoplasm results in a more complex pattern of diffusion. It is also necessary to clarify whether such changes affect local biochemical signaling events and molecular localizations. The number of molecules per spine could influence the magnitude of functional changes (Okabe, 2007; Ribrault et al., 2011). Furthermore, the changes in diffusion induced by alterations in spine structure will affect the stability of the structure. This will subsequently change the molecule’s diffusivity. Thus, it will be interesting to investigate whether this type of mutual relationship exists within spines.New imaging techniques will help to answer these questions. By applying fast 3-D SPT to intra-spine measurements, it will be possible to investigate the spatial heterogeneity of diffusion in single spines of living neurons in detail (Hou et al., 2020; Xiang et al., 2020). STED-FCS/fluorescence cross-correlation spectroscopy can also detect changes in intermolecular interactions (Lanzanò et al., 2017). In addition to the development of new measurement techniques, molecular dynamics simulations based on experimental data will become increasingly important in the future (Okabe, 2020a; Reshetniak et al., 2020a; Vasan et al., 2020). Spine morphology and intra-spine structures, which affect diffusion, are closely related. Thus, it is difficult to investigate the effect of one without changing the other experimentally. Molecular dynamics simulation is a useful tool to examine how molecular motion is adjusted by combining elements that are difficult to verify experimentally (Bell et al., 2019). Furthermore, the shape of spines and intra-spine components, such as the actin cytoskeleton, which are the structural basis of spines, differ from spine to spine. Here, a combination of quantitative measurements and simulations based on experimental data will help us to understand molecular events more quantitatively.Although we reviewed work using fluorescence microscopy, details of spine morphology and intra-spine structures have also been revealed by EM at the nanoscale (Bourne and Harris, 2012; Tao et al., 2018). However, it is difficult to observe specific molecular localizations with EM. On the other hand, super-resolution microscopy is suitable for obtaining a nanoscale picture of molecular positions within spines. Yet, it is still difficult to observe dense structures such as actin polymers (Kommaddi et al., 2018). Therefore, in the future, it will be essential to combine the advantages of each technique, observing internal structures at the nanoscale using EM and measuring molecular localization with super-resolution microscopy (CLEM; Taraska, 2019; Hoffman et al., 2020). Of course, dynamic intracellular structures such as lipid rafts and biomolecular condensates are also likely to affect molecular mobility (Sezgin et al., 2017; Chen et al., 2020). Thus, it will be key to overlay molecular mobilities from living cells over the static structural information of CLEM. We believe that combinations of multiple imaging modalities, along with modeling, will allow for a more in-depth understanding of synapses at the molecular level. These data will reveal how the elaborate architecture, density, and compartmentalization of subcellular components influence the highly tuned, dynamic, and changeable actions of synapses in the brain.  相似文献   

13.
14.
Teleost fishes are the most species-rich clade of vertebrates and feature an overwhelming diversity of sex-determining mechanisms, classically grouped into environmental and genetic systems. Here, we review the recent findings in the field of sex determination in fish. In the past few years, several new master regulators of sex determination and other factors involved in sexual development have been discovered in teleosts. These data point toward a greater genetic plasticity in generating the male and female sex than previously appreciated and implicate novel gene pathways in the initial regulation of the sexual fate. Overall, it seems that sex determination in fish does not resort to a single genetic cascade but is rather regulated along a continuum of environmental and heritable factors.IN contrast to mammals and birds, cold-blooded vertebrates, and among them teleost fishes in particular, show a variety of strategies for sexual reproduction (Figure 1), ranging from unisexuality (all-female species) to hermaphroditism (sequential, serial, and simultaneous, including outcrossing and selfing species) to gonochorism (two separate sexes at all life stages). The underlying phenotypes are regulated by a variety of sex determination (SD) mechanisms that have classically been divided into two main categories: genetic sex determination (GSD) and environmental sex determination (ESD) (Figure 2).Open in a separate windowFigure 1Reproductive strategies in fish. Fish can be grouped according to their reproductive strategy into unisexuals, hermaphrodites, and gonochorists. Further subdivisions of these three categories are shown with pictures of species exemplifying the strategies. Fish images: Amphiprion clarkii courtesy of Sara Mae Stieb; Hypoplectrus nigricans courtesy of Oscar Puebla; Scarus ferrugineus courtesy of Moritz Muschick; Astatotilapia burtoni courtesy of Anya Theis; Poecilia formosa and Kryptolebias marmoratus courtesy of Manfred Schartl; Trimma sp. courtesy of Rick Winterbottom [serial hermaphroditism has been described in several species of the genus Trimma (Kuwamura and Nakashima 1998; Sakurai et al. 2009; and references therein)].Open in a separate windowFigure 2Sex-determining mechanisms in fish. Sex-determining systems in fish have been broadly classified into environmental and genetic sex determination. For both classes, the currently described subsystems are shown.Environmental factors impacting sex determination in fish are water pH, oxygen concentration, growth rate, density, social state, and, most commonly, temperature (for a detailed review on ESD see, e.g., Baroiller et al. 2009b and Stelkens and Wedekind 2010). As indicated in Figure 2, GSD systems in fish compose a variety of different mechanisms and have been reviewed in detail elsewhere (e.g., Devlin and Nagahama 2002; Volff et al. 2007).The GSD systems that have received the most scientific attention so far are those involving sex chromosomes, which either may be distinguishable cytologically (heteromorphic) or appear identical (homomorphic). In both cases, one sex is heterogametic (possessing two different sex chromosomes and hence producing two types of gametes) and the other one homogametic (a genotype with two copies of the same sex chromosome, producing only one type of gamete). A male-heterogametic system is called an XX-XY system, and female-heterogametic systems are denoted as ZZ-ZW. Both types of heterogamety exist in teleosts and are even found side by side in closely related species [e.g., tilapias (Cnaani et al. 2008), ricefishes (Takehana et al. 2008), or sticklebacks (Ross et al. 2009)]; for more details on the phylogenetic distribution of GSD mechanisms in teleost fish, see Mank et al. (2006). Note that sex chromosomes in fish are mostly homomorphic and not differentiated (Ohno 1974), which is in contrast to the degenerated Y and W chromosomes in mammals (Graves 2006) and birds (Takagi and Sasaki 1974), respectively. This is one possible explanation for the viable combination of different sex chromosomal systems within a single species or population of fish (Parnell and Streelman 2013) and could be a mechanistic reason why sex chromosome turnovers occur easily and frequently in this group (Mank and Avise 2009). Additionally, fish can have more complex sex chromosomal systems involving more than one chromosome pair (see Figure 2). Even within a single fish species, more than two sex chromosomes may occur at the same time, or more than two types of sex chromosomes may co-exist in the same species (Schultheis et al. 2006; Cioffi et al. 2013), which can sometimes be due to chromosome fusions (Kitano and Peichel 2012).Detailed insights on the gene level for GSD/sex chromosomal systems are currently available for only a limited number of fish species, and all but one of these cases involve a rather simple genetic system with male heterogamety and one major sex determiner (see below). The only exception is the widely used model species zebrafish (Danio rerio), which has a polyfactorial SD system implicating four different chromosomes (chromosomes 3, 4, 5, and 16) (Bradley et al. 2011; Anderson et al. 2012) and also environmental cues (Shang et al. 2006).In this review, we focus on newly described genetic sex-determining systems and possible mechanisms allowing their emergence in fishes, which are the most successful group of vertebrates with ∼30,000 species.  相似文献   

15.
Regulation of autophagy in neurons remains unclear. In this issue, Kulkarni et al. (2021. J. Cell Biol. https://doi.org/10.1083/jcb.202002084) show with elegant live imaging that in dendrites, but not in axons, autophagosome motility and function is regulated by synaptic activity.

Macroautophagy is a type of autophagy that refers to the capacity to form double membrane compartments called autophagosomes that engulf large protein aggregates and defective organelles. Autophagosomes fuse with lysosomes, forming degradative autolysosomes (1). Autophagosome formation depends on the conjugation of LC3-I (cytosolic) to phosphatidylethanolamine, generating LC3-II, which remains bound to autolysosomes (1). In neurons, inactivation of autophagy genes impacts neurodevelopment, axon growth and guidance, synapse formation and pruning, ultimately leading to neurodegeneration. Particularly, in motor neurons and cerebellum Purkinje cells, autophagy gene knockout leads to the accumulation of intracellular protein aggregates and degeneration, impacting movement coordination (1). Interestingly, stimulation of memory up-regulates autophagy, and while reducing autophagy reduces memory, activating it has the opposite effect on memory (2). What triggers macroautophagy in neurons remains unclear. In this issue, Kulkarni et al. test whether synaptic activity regulates autophagy and detail the impact of synaptic activity on autophagosome motility (3).Kulkarni et al. used multiple strategies to manipulate synaptic activity. They stimulated synaptic activity by depolarizing neurons with high potassium, treating them with a cocktail of antagonists of voltage-gated potassium channels and inhibitory gamma-aminobutyric A receptors, and using uncaging of the excitatory neurotransmitter glutamate. To inhibit synaptic activity, the researchers treated neurons with antagonists of excitatory α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid and N-methyl-D-aspartate receptors (4). To image autophagosomes and autolysosomes (here globally termed autophagic vacuoles [AVs]) in live neurons, the authors expressed LC3 tagged with fluorescent proteins. They elegantly imaged the same neuronal compartment before and after depolarization, or under basal, increased, or reduced synaptic activity, and used kymograph analysis (via Kymoanalyser; 5) to quantify the mean speeds of AVs in both dendrites and axons. An increase in intracellular calcium measured with a genetically encoded calcium sensor, GCAMP3, indicated synaptic activity. Kulkarni et al. observed that, in dendrites, AVs stop with synaptic activity and move with synaptic inhibition (Fig. 1). This AV movement change was swift and unaltered by co-culture with astrocytes, and reversible. One key finding is that this change in AV movement occurred in dendrites, but not in axons. Interestingly, AVs stopped at or near synapses, which were identified with PSD-95-GFP.Open in a separate windowFigure 1.In dendrites, AVs stop at synapses upon synaptic activity.The authors further characterized the AVs in terms of acidity (lysotracker labelling of acidic organelles) and of degradative capacity (DQ-BSA fluorescence accumulation upon lysosomal degradation). Lysotracker motility changed similarly with synaptic activity. Interestingly, the lysotracker density increased with synaptic stimulation. The higher number of acidic organelles (likely autolysosomes) indicated increased autophagy or acidification with synaptic activity, which could underlie increased degradative activity. Indeed, about half of the LC3-positive AVs were degradative in dendrites, while in axons there was virtually no degradative AV, supporting the requirement for transport to the soma for degradation of autophagic cargo (6). Finally, Kulkarni et al. show that degradative AVs increase with synaptic activity, correlating with the reduced motility of LC3-positive AVs.An intriguing observation is that the autophagic vacuoles identified by LC3-mCherry were virtually all positive for LAMP1, a marker of late endosomes and lysosomes, indicating that dendrites mainly contain autolysosomes and no or very few autophagosomes (LC3-positive and LAMP1-negative) and late endosomes/lysosomes (LC3-negative and LAMP1-positive). One is left wondering if it results from LC3 overexpression and overflooding to interconnected organelles. An alternative possibility is that LC3 may not always label autophagosomes, in which case complementary electron microscopy is necessary for confirmation. Where are dendritic autolysosomes formed? In axons, a fraction of the LC3 autophagic vacuoles was LAMP1 negative, and the formation of autophagosomes at axon terminals has been well documented (7). Thus, do autophagosomes form in axons, fuse with LAMP1-positive late endosomes/lysosomes, and only after are they transported to dendrites? Alternatively, autophagosomes may form in dendrites and fuse with late endosomes/lysosomes, preventing their detection unless fusion is inhibited (8).Another interesting observation concerns the similar change in the motility of early endosomes, identified by Rab5, an early endosome GTPase, with synaptic activity. Other organelles, post-ER vesicles (4), and proteasomes (9) similarly display a change in motility in dendrites upon synaptic activity. In contrast, mitochondria stop moving in axons with synaptic activity (10). The significance of this arrest of several dendritic organelles with synaptic activity is an attractive area for research.Neuronal autophagy dysfunction is implicated in many neurodegenerative diseases (1). At least early in the disease, increasing autophagy improves neuronal function and synapse activity (1). Genetic risk factors include lysosomal proteins, whose defective function leads to the accumulation of nondegraded autophagic vacuoles and recapitulate neurodegenerative phenotypes (11). Lysosomal dysfunction is a mechanism of cellular aging. Moreover, synapses become dysfunctional with aging and lost in neurodegenerative diseases (12). Based on this study, synapse dysfunction and thus reduced synaptic activity could increase AV motility and reduce acidification and the degradative capacity of autolysosomes. Similarly, neuronal overexcitability, as in early Alzheimer''s disease patients with seizures, could cause excessive AV motility and degradative activity.What is the mechanism that stops AV movement? Do early endosomes, secretory vesicles, or proteasomes change motility using similar mechanisms? For post-ER vesicles, the CAMKII dependent phosphorylation of the microtubule motor Kif17 was sufficient to arrest movement (4). Alternatively, could it be the actin cytoskeleton that forms patches in the dendritic shaft at the base of postsynaptic glutamatergic synapses to halt microtubule-dependent transport of organelles (13)? More work is needed to tackle these questions and define the cell biological mechanisms by which synaptic activity controls AV function and dynamics in different neuronal compartments. Understanding the mechanisms underlying the regulation of autophagy and autophagosome maturation and degradation provides an exciting opportunity for therapeutic development in neurodegenerative diseases.  相似文献   

16.
Silencing of the spindle assembly checkpoint involves two protein phosphatases, PP1 and PP2A-B56, that are thought to extinguish checkpoint signaling through dephosphorylation of a checkpoint scaffold at kinetochores. In this issue, Cordeiro et al. (2020. J. Cell Biol. https://doi.org/10.1083/jcb.202002020) now show that a critical function of these phosphatases in checkpoint silencing is removal of Polo kinase at kinetochores, which would otherwise autonomously sustain the checkpoint.

The main goal of mitosis is to accurately segregate chromosomes, such that each daughter cell inherits a full complement of genetic information. To accomplish this delicate task, once each chromosome is faithfully duplicated through DNA replication, its identical sister chromatids must attach to spindle microtubules coming from opposite spindle poles through a process known as chromosome biorientation. Kinetochores are proteinaceous assemblies that reside at the centromeric region of chromosomes and are key to this process by capturing spindle microtubules (1). Microtubule capture, however, is inherently error prone, and several cycles of attachment/detachment are often required before chromosomes achieve biorientation. Obviously, chromosome segregation without error correction would be highly detrimental, leading to unbalanced chromosome numbers, referred to as aneuploidies, which are hallmarks of cancer and genetic diseases. Luckily, eukaryotic cells not only possess an error-correction machinery deputed to rectify faulty attachments (2), but they also have a safeguard device, called the spindle assembly checkpoint (SAC), that temporarily halts cells in mitosis to provide them with the necessary time window to fix the errors. SAC signaling fires at unattached kinetochores, which are continuously generated during error correction, and is extinguished once all chromosomes are bioriented, thus resuming mitotic progression and chromosome segregation (3).Prevailing models posit that a key trigger of SAC signaling is the phosphorylation of the kinetochore scaffold KNL1 by the SAC kinase MPS1. This creates a phospho-docking site at the MELT repeats (amino acid consensus Met-Glu-Leu-Thr) of KNL1 that recruits the heterotetrameric SAC complex BUB1:BUB3:BUB3:BUBR1 (referred to as BUB complex; Fig. 1), which in turn attracts to the kinetochore other SAC proteins that collectively prevent mitotic progression (3).Open in a separate windowFigure 1.The interplay of SAC kinases and phosphatases at kinetochores. When SAC is activated at an unattached kinetochore (SAC on), MPS1 phosphorylates the kinetochore scaffold KNL1, thereby recruiting the BUB complex. Contextually, CDK1-dependent phosphorylation of BUB1 and BUBR1 generates phospho-docking sites for recruitment of the Polo kinase PLK1, which on one side sustains KNL1 phosphorylation and on the other stimulates BUBR1 binding to PP2A-B56. The latter, in turn, counteracts PLK1 local activity by dislodging PLK1 from the kinetochore. During SAC silencing, local activity of MPS1 is shut off. Additionally, the PP1 phosphatase binds to KNL1 and, together with PP2A-B56, further evicts PLK1 from the kinetochore, possibly through dephosphorylation of its phospho-docking sites in BUB1 and BUBR1. This leads to KNL1 dephosphorylation and displacement of the BUB complex, thus extinguishing SAC signaling (SAC off). Whether PP1 and PP2A-B56, as opposed to other phosphatases, contribute directly to KNL1 dephosphorylation remains an open question.The protein phosphatases PP1 and PP2A-B56 are recruited to kinetochores through binding to KNL1 and BUBR1, respectively, and are thought to silence the SAC through dephosphorylation of the MELT repeats of KNL1, thus antagonizing MPS1 activity (Fig. 1). Additional mechanisms, such as MPS1 inhibition and stripping of SAC components from kinetochores, have been proposed to contribute to obliterate SAC signaling upon chromosome biorientation (4).In this issue, compelling evidence from Cordeiro et al. challenges the current view by showing that rather than dephosphorylating KNL1, PP1 and PP2A-B56 actually silence the SAC by down-regulating the activity of Polo kinase (PLK1 in human cells) at kinetochores (5). Polo kinase and MPS1 share a common substrate preference and both can phosphorylate the MELT repeats of KNL1. Additionally, Polo cooperates with MPS1 in SAC signaling in various species, while in organisms where MPS1 is absent, like nematodes, Polo functionally replaces MPS1 (6).Consistent with previous results (7, 8), Cordeiro et al. show that when kinetochore phosphatases are dampened, PLK1 levels increase at kinetochores through an unknown mechanism, which might involve dephosphorylation of the phosphoepitopes in the Polo-binding motifs generated on the BUB complex by CDK1 (BUB1-pT609 and BUBR1-pT620; 9, 10, 11). This implies that when PP1 and PP2A-B56 are low at kinetochores, PLK1 can amplify SAC signaling through a positive feedback loop by boosting KNL1 phosphorylation independently of Mps1, thereby recruiting the BUB complex and, in turn, increasing amounts of PLK1 (Fig. 1). In agreement with this view, in a sensitized setup where kinetochore phosphatases are crippled along with MPS1, concomitant inhibition of PLK1 is sufficient to bring about KNL1 dephosphorylation and restore SAC signaling. These data led the authors to the provocative conclusion that the primary role of PP1 and PP2A-B56 in SAC silencing is to harness PLK1 activity. This new model is appealing not only because it highlights a novel function for PP1 and PP2A-B56 in SAC silencing, but also because it explains the modest effects that are commonly observed on SAC signaling upon PLK1 inhibition alone. Indeed, kinetochore phosphatases, and primarily PP2A-B56 (12), are already partially active in a SAC-induced mitotic arrest (e.g., upon microtubule depolymerization), as shown here by the increased KNL1 phosphorylation upon their inactivation.Interestingly, sequence alignment of BUB1 and BUBR1 homologues across the phylogenetic tree reveals that, in metazoans, putative Polo-binding motifs are usually located in the vicinity of hypothetical PP2A-B56–binding motifs, suggesting that they coevolved. The physical proximity of Polo-binding and PP2A-B56–binding motifs in BUB1 and BUBR1 could position the Polo-binding motifs in a favorable arrangement for their PP2A-B56–driven dephosphorylation and, as a consequence, PLK1 clearance from kinetochores (Fig. 1).The data by Cordeiro et al. represent a paradigm shift in our understanding of SAC silencing for two main reasons. First, consistent with published data (13, 14), PP1 and PP2A-B56 might be involved in this process primarily by inactivating upstream SAC kinases (MPS1 and PLK1), rather than dephosphorylating their substrates. Second, since PLK1 is partially displaced from kinetochores by the above phosphatases already during a SAC arrest, MPS1 inactivation might be the main trigger of SAC silencing. Several mechanisms have been proposed to attenuate MPS1 activity once the SAC is satisfied, such as MPS1 displacement from kinetochores (6) and dephosphorylation of MPS1 in its activation loop (13, 14).The new model raises a burning question: If PP1 and PP2A do not dephosphorylate KNL1 at MELT repeats, what does? Other phosphatases, whose identity remains elusive, could be involved in KNL1 dephosphorylation. Alternatively, phosphorylated KNL1 might be actively turned over at kinetochores. Nevertheless, at present, the involvement of PP1 and PP2A-B56 in KNL1 dephosphorylation cannot be ruled out, as complete inhibition of kinetochore phosphatases in the experimental setup used here is likely very challenging. Further investigations will be required to solve this central issue.Another important question that deserves further scrutiny is, how exactly do PP1 and PP2A-B56 inhibit PLK1 activity at kinetochores? Cordeiro et al. propose that they could dephosphorylate the Polo-binding motifs in BUB1/BUBR1. Alternatively, the close proximity of PP2A- and Polo-binding motifs in metazoan BUBR1 homologues could make the association of PLK1 and PP2A with BUBR1 mutually exclusive.Finally, and most importantly, what is the physiological meaning of the complex interplay between SAC kinases and phosphatases described here? A crucial function of PLK1 bound to the BUB complex in human cells is to stabilize kinetochore-microtubule attachments in prometaphase by recruiting PP2A-B56 through phosphorylation of the PP2A-B56-binding motif in BUBR1 (15). In turn, eviction of PLK1 from kinetochores by PP2A-B56 will have two major outputs: (i) maintain microtubule dynamics at bioriented chromosomes (8) and (ii) stimulate binding of PP1 to KNL1, which primes the system for SAC silencing (16). As soon as MPS1 levels drop at kinetochores and/or other phosphatases intervene to dephosphorylate KNL1, SAC signaling is finally extinguished. The development of fluorescence-based biosensors combined with mathematical modeling will certainly provide in the future further mechanistic insights into such intricate network.  相似文献   

17.
The turnover of actin filament networks in cells has long been considered to reflect the treadmilling behavior of pure actin filaments in vitro, where only the pointed ends depolymerize. Newly discovered molecular mechanisms challenge this notion, as they provide evidence of situations in which growing and depolymerizing barbed ends coexist.

IntroductionIn cells, actin assembles into filament networks with diverse architectures and lifetimes, playing key roles in functions such as endocytosis, cell motility, and cell division. These filament networks are maintained and renewed by actin turnover, which implies that assembly and disassembly must take place simultaneously and in a controlled manner within the networks. Each actin filament end has the ability to either grow or shrink, depending on the concentration of actin and regulatory proteins, but pure actin treadmills at steady state: ATP-actin is added at the barbed end at a rate matching the departure of ADP-actin from the pointed end, and ATP hydrolysis takes place within the filament. This hallmark feature of actin dynamics has been known for decades (Wegner, 1976) and has been generalized to the cell context, in which it is commonly assumed that actin polymerization takes place at the barbed end, while depolymerization takes place only at the pointed end (whether it be the ends of filaments within the network or the ends of fragments that have detached from it). This notion is reinforced by the fact that the cytoplasm contains high concentrations of monomeric actin (G-actin) in complex with profilin (Funk et al., 2019), which is unable to bind to pointed ends and should drive the elongation of all noncapped barbed ends.Recently, however, in vitro studies have identified two seemingly independent mechanisms in which, in the presence of profilin-actin, filament barbed ends alternate between phases of growth and depolymerization. This behavior, referred to as “dynamic instability,” is widely observed for microtubules but was unexpected for actin filaments. It suggests that cells could use barbed ends for both elongation and disassembly.Driving the depolymerization of barbed ends with cofilin side-decorationProteins of the actin depolymerizing factor (ADF)/cofilin family (henceforth cofilin) are composed of a single ADF-homology (ADF-H) domain and are mostly known for their actin filament–severing activity (De La Cruz, 2009). Cofilin binds cooperatively to the sides of actin filaments, forming clusters where the conformation of the filament is locally altered, leading to its severing at cofilin cluster boundaries. In addition, the barbed ends of cofilin-decorated filaments steadily depolymerize, despite the presence of G-actin and profilin-actin (Fig. 1 A) and even capping protein (CP) in solution (Wioland et al., 2017, 2019). This unexpected result likely originates from the conformational change of actin subunits at the barbed end, induced by cofilin side-binding. As a consequence, filaments exposed to G-actin (with or without profilin), CP, and cofilin alternate between phases of barbed-end elongation and barbed-end depolymerization. In these conditions, actin filament barbed ends thus exhibit a form of dynamic instability.Open in a separate windowFigure 1.Two mechanisms that give rise to barbed-end depolymerization in elongation-promoting conditions. (A) When a cofilin side-decorated region reaches the barbed end, adding a new actin or profilin-actin becomes very difficult, and the barbed end depolymerizes. Not represented: Capping by CP can lead to depolymerization, as it allows the cofilin cluster to reach the barbed end, which then has a much weaker affinity for CP and steadily depolymerizes. Also, severing events occur at cofilin cluster boundaries, creating new barbed ends, either bare or cofilin-decorated. (B) Twinfilin binds to the barbed end, preventing its elongation and causing its depolymerization. Whether twinfilin remains processively attached to the depolymerizing barbed end or departs with the actin subunits is still unknown. Twinfilin has no impact on the elongation of mDia1-bearing barbed ends.Driving the depolymerization of barbed ends with twinfilin end-targetingTwinfilin has two ADF-H domains, but unlike cofilin, it binds poorly to the sides of actin filaments. Rather, twinfilin appears to mainly sequester ADP-actin monomers and target the barbed end to modulate its elongation and capping. Recent in vitro studies have shown that the interaction of twinfilin with actin filament barbed ends could drive their depolymerization, even in the presence of G-actin and profilin-actin (Johnston et al., 2015; Hakala et al., 2021; Shekhar et al., 2021). Very interestingly, the processive barbed-end elongator formin mDia1 is able to protect barbed ends from twinfilin, allowing them to sustain elongation (Shekhar et al., 2021). This leads to a situation in which, as filaments are exposed to profilin-actin and twinfilin, mDia1-bearing barbed ends elongate while bare barbed ends depolymerize (Fig. 1 B). It is safe to assume that, if filaments were continuously exposed to this protein mix including formin in solution, they would alternate between phases of growth and shrinkage over time, as formins come on and fall off the barbed end. This mix of proteins would therefore constitute another situation causing actin filament dynamic instability.From actin treadmilling to dynamic instability, in cells?This newly identified versatile behavior of actin filaments is reminiscent of microtubules. While dynamic instability is the hallmark behavior of microtubules, they can also be made to treadmill steadily by adding 4 microtubule-associated proteins (Arpağ et al., 2020). In cells, both microtubule dynamic instability and treadmilling have been clearly observed (Wittmann et al., 2003). In contrast, the disassembly of single actin filaments, either embedded in a network or severed from it, has not yet been directly observed in cells. Despite insights from techniques such as single-molecule speckle microscopy, it is still unclear from which end actin filaments depolymerize, even in networks that appear to globally treadmill, such as the lamellipodium. Pointed end depolymerization alone cannot account for what is observed in cells (Miyoshi et al., 2006) and alternative mechanisms have been proposed, including brutal filament-to-monomer transitions occurring in bursts, driven by cofilin, coronin, and Aip1 (Brieher, 2013; Tang et al., 2020).In cells, the high amounts of available G-actin (tens of micromolars; Funk et al., 2019) should limit barbed-end depolymerization. Based on the reported on-rate for ATP–G-actin at the barbed ends of cofilin-decorated filaments (Wioland et al., 2017, 2019), we can estimate that these barbed ends, under such conditions, would depolymerize for tens of seconds before being “rescued,” which is enough to remove tens of subunits from each filament. In contrast, twinfilin concentrations similar to those of G-actin appear necessary to drive barbed-end depolymerization (Hakala et al., 2021; Shekhar et al., 2021). As proteomics studies in HeLa cells report that twinfilin is 50-fold less abundant than actin, this may be difficult to achieve in cells (Bekker-Jensen et al., 2017). However, future studies may uncover proteins, or posttranslational modifications of actin, that enhance the ability of twinfilin to drive barbed-end depolymerization in the presence of high concentrations of profilin-actin.Molecular insights and possible synergiesWhile cofilin and twinfilin both interact with actin via ADF-H domains, they appear to drive barbed-end depolymerization through different mechanisms: twinfilin by directly targeting the barbed end, and cofilin by decorating the filament sides, thereby changing the conformation of the filament and putting its barbed end in a depolymerization-prone state.The two mechanisms, nonetheless, share clear similarities. For instance, cofilin side-binding and twinfilin end-targeting both slow down ADP-actin barbed-end depolymerization, compared with bare ADP-actin filaments (Wioland et al., 2017; Hakala et al., 2021; Shekhar et al., 2021). Strikingly, a crystal structure of the actin/twinfilin/CP complex indicates that the actin conformational change induced by twinfilin binding at the barbed end is similar to that induced by cofilin decorating the sides (Mwangangi et al., 2021). It is thus possible that the dynamic instability of actin filament barbed ends reflects the same conformation changes, triggered either by cofilin side-decoration or twinfilin end-targeting.In addition to decorating the filament sides, cofilin targets ADP-actin barbed ends. Unlike twinfilin, the direct interaction of cofilin with the barbed end cannot cause its depolymerization in the presence of ATP-actin monomers. Indeed, cofilin end-targeting accelerates the depolymerization of ADP-actin barbed ends in the absence of G-actin, but cofilin does not appear to interact with growing ATP-actin barbed ends (Wioland et al., 2017). This is in stark contrast with twinfilin end-targeting, which slows down ADP-actin depolymerization and accelerates ADP–Pi-actin depolymerization (Shekhar et al., 2021). These different behaviors regarding the nucleotide state of actin are intriguing and should be investigated further.Cofilin thus needs to decorate the filament sides in order to have an impact on barbed-end dynamics in elongation-promoting conditions. However, it is unknown whether cofilin side-decoration extends all the way to the terminal subunits and occupies sites that twinfilin would target. Thus, it is unclear whether cofilin and twinfilin would compete or synergize to drive barbed-end depolymerization.Synergies with other proteins are also worth further investigation, CP being an interesting candidate. Cofilin side-decoration drastically decreases the barbed-end affinity for CP, and capped filaments are thereby an efficient intermediate to turn growing barbed ends into depolymerizing barbed ends (Wioland et al., 2017). Twinfilin interacts with CP and the barbed end to enhance uncapping (Hakala et al., 2021; Mwangangi et al., 2021). Since CP can bind mDia1-bearing barbed ends and displace mDia1 (Bombardier et al., 2015; Shekhar et al., 2015), perhaps CP can also contribute to turn growing, mDia1-bearing barbed ends into depolymerizing barbed ends, by removing mDia1 from barbed ends and subsequently getting displaced from the barbed end by twinfilin.Finally, it is worth noting that profilin, which does not contain an ADF-H domain, also interacts with the barbed face of G-actin and with the barbed end of the filament. When profilin is in sufficient excess, it is able to promote barbed-end depolymerization in the presence of ATP–G-actin (Pernier et al., 2016). Unlike twinfilin, its depolymerization-promoting activity is not prevented by formin mDia1, and it thus does not lead to dynamic instability (bare and mDia1-bearing barbed ends all either grow or depolymerize). The coexistence of growing, mDia1-bearing barbed ends and depolymerizing, twinfilin-targeted barbed ends (Fig. 1 B) was observed in the presence of profilin (Shekhar et al., 2021), but profilin actually may not be required. Future studies should determine the exact role of profilin in this mechanism.ConclusionThe extent to which barbed-end dynamic instability contributes to actin turnover in cells is not known, but possible molecular mechanisms have now been identified. They should change the way we envision actin network dynamics, as we must now consider the possibility that cells also exploit the barbed end for disassembly. More work is needed to further document these mechanisms, but the idea of a “generalized treadmilling” has now been contradicted at its source: in vitro experiments.  相似文献   

18.
We have long known that lipids traffic between cellular membranes via vesicles but have only recently appreciated the role of nonvesicular lipid transport. Nonvesicular transport can be high volume, supporting biogenesis of rapidly expanding membranes, or more targeted and precise, allowing cells to rapidly alter levels of specific lipids in membranes. Most such transport probably occurs at membrane contact sites, where organelles are closely apposed, and requires lipid transport proteins (LTPs), which solubilize lipids to shield them from the aqueous phase during their transport between membranes. Some LTPs are cup like and shuttle lipid monomers between membranes. Others form conduits allowing lipid flow between membranes. This review describes what we know about nonvesicular lipid transfer mechanisms while also identifying many remaining unknowns: How do LTPs facilitate lipid movement from and into membranes, do LTPs require accessory proteins for efficient transfer in vivo, and how is directionality of transport determined?

IntroductionEukaryotic cells contain diverse membranes, each with a characteristic and carefully regulated protein and lipid content. Most membrane proteins are first inserted into the ER and then traffic among cellular compartments in vesicles (Mellman and Warren, 2000). The same is true of lipids—most are synthesized in the ER and are then exchanged between organelles in vesicles (Vance, 2015). Lipids, however, are also moved between organelles by nonvesicular pathways. This type of transport has a number of functions. One is the bulk transfer of lipids sufficient to sustain organelle biogenesis. For example, mitochondria membrane biogenesis requires the import of most lipids (Acoba et al., 2020). Nonvesicular lipid transport also allows cells to change the lipid composition more rapidly and precisely than is possible by vesicular trafficking or during stress conditions when vesicular trafficking is compromised. It can be used to enrich or deplete membranes of particular lipids, either to modulate the lipid composition of a membrane or to regulate levels of signaling lipids.Nonvesicular lipid exchange within cells could, in theory, occur by two mechanisms. One is lipid diffusion between membranes following hemifusion, where the outer leaflets of two bilayers merge. While hemifusion is thought to occur transiently during membrane fusion, there is no evidence that this mechanism is used to move lipids between cellular compartments that do not fuse. The second type of mechanism is transport by proteins known as lipid transport proteins (LTPs). These proteins have the ability to move lipids between membranes via hydrophobic cavities that shield the lipids from the aqueous environment during transport. There are many families of LTPs, with most eukaryotic cells expressing multiple members of each family (Chiapparino et al., 2016; Wong et al., 2019). Although most of the known transporters function as shuttles that carry single lipid molecules between membranes, others serve as bridges that allow lipids to flow between membranes (Li et al., 2020). Mutations in some LTPs are known to cause diseases (Table 1).Table 1.Diseases caused by mutations in LTPs
LTPDiseaseReferences
α-Tocopherol transfer proteinAtaxia with isolated vitamin E deficiency (OMIM 277460)Ouahchi et al., 1995
Microsomal triglyceride transfer proteinAbetalipoproteinemia (OMIM 200100)Shoulders et al., 1993
Niemann-Pick C1 proteinNiemann-Pick disease type C1 (OMIM 257220)Carstea et al., 1997
Niemann-Pick C2 proteinNiemann-Pick disease type C2 (OMIM 607625)Naureckiene et al., 2000
Steroid acute regulatory proteinLipoid congenital adrenal hyperplasia (OMIM 201710)Lin et al., 1995
Vps13AChorea-acanthocytosis (OMIM 200150)Rampoldi et al., 2001
Vps13BCohen syndrome (OMIM 216550)Kolehmainen et al., 2003
Vps13CEarly onset Parkinson’s disease (OMIM 616840)Lesage et al., 2016
Vps13DSpastic ataxia (OMIM 607317)Gauthier et al., 2018; Seong et al., 2018
Open in a separate windowOMIM, Online Mendelian Inheritance in Man number.Most LTPs operate at membrane contact sites (MCSs), regions where two organelles are closely apposed. Localization at these sites could serve to speed transport by reducing the distance that LTPs must diffuse as they shuttle lipids between membranes. However, some LTPs do not operate at MCSs, (e.g., STARD4 [Mesmin et al., 2011] and ORP2 [Wang et al., 2019]), and it has been argued that LTP diffusion is not the rate-limiting step of lipid transport (Dittman and Menon, 2017). The enrichment LTPs at MCSs could have functions other than increasing the transport rate, such as facilitating LTP interaction with proteins that modulate lipid transport or restricting lipid exchange to a specific pair of organelles.This review focuses on our emerging understanding of how nonvesicular lipid transport occurs and identifies important challenges and unanswered questions in the field. It begins by summarizing what we know about the rates and volumes of lipid exchange in cells and the general structural features of LTPs. We then discuss open questions about mechanisms of lipid transport and what drives lipid transport.Volume and rates of nonvesicular lipid transport in cellsCells have considerable capacity to move some types of lipids between organelles by nonvesicular pathways. This was suggested by early studies on lipid transport from the ER, where most lipids are synthesized, to the plasma membrane. These studies concluded that newly synthesized glycerophospholipids, cholesterol, and glucosylceramides are transferred to the plasma membrane by nonvesicular mechanisms, because lipid transport was not blocked by inhibiting vesicular trafficking and occurred at rates that were too rapid to be explained by vesicular trafficking (Kaplan and Simoni, 1985a; Kaplan and Simoni, 1985b; Sleight and Pagano, 1983; Vance et al., 1991; Warnock et al., 1994). There is also evidence that exogenous sterols, which first enter cells by incorporating into the plasma membrane, are rapidly exchanged between the plasma membrane and other organelles by nonvesicular transport pathways. When the naturally fluorescent sterol, dehydroergosterol (DHE), is added to cells, it equilibrates between the plasma membrane and endocytic recycling compartments in minutes, and it has been estimated that there are ∼106 molecules of DHE exchanged between these organelles per second in CHO cells (Maxfield and Mondal, 2006). This is a remarkable volume of transport given that there are ∼3 × 108 cholesterol molecules in the plasma membrane of this cell type, suggesting that these cells have the capacity to exchange all the cholesterol in the plasma membrane in ∼5 min. This volume of DHE transport is more than can be explained by vesicular trafficking.There must also be a substantial amount of nonvesicular lipid transport to such organelles as mitochondria and chloroplasts, which are largely disconnected from vesicular trafficking pathways. This has been termed bulk transport, since the function is to provide sufficient lipid to support membrane expansion (Fig. 1 A). To put in perspective the volume of nonvesicular phospholipid transport required to sustain membrane biogenesis, consider mitochondrial membrane biogenesis in the yeast Saccharomyces cerevisiae. It has been estimated that mitochondrial biogenesis requires the transport of ∼20,000 phospholipids per second when this yeast is growing at top speed (doubling every 2 h; Petrungaro and Kornmann, 2019). A similarly high rate of phospholipid transport is required for the maturation of nascent autophagosomes, which can occur in minutes and has been estimated to require the transport of ∼108 lipids per cell (Melia et al., 2020).Open in a separate windowFigure 1.Functions of nonvesicular lipid transport. (A) High-volume lipid transport required for membrane expansion (bulk transport). Shown for growth of an autophagosome and also necessary for biogenesis of mitochondria and chloroplasts. (B) Types of lower-volume lipid transport. Representative examples of three functions are shown. One is to support lipid-based signaling (left). The protein Nir2 transfers PI from the ER to the plasma membrane, where it is converted to PI4P by the enzyme phosphatidylinositol 4-kinase-α (PI4KA). PI4P can removed from the plasma membrane by ORP5 or ORP8, which bring it to the ER, where it is hydrolyzed by the phosphatidylinositide phosphatase, Sac1. Lipid transport can also regulate the levels of a specific lipid. For example, OSBP uses counter-exchange transport to enrich cholesterol in the Golgi membrane (center). Lipid transport can also support membrane domain formation. For example, Lam6/Ltc1 brings the sterol, ergosterol, to the vacuole in S. cerevisiae and supports membrane domain formation there (right; domain in red).In addition to bulk transport to support membrane expansion, cells use lower-volume transport in three ways (Fig. 1 B). One way is to support phosphoinositide (PIP) signaling and other signaling that uses lipid messengers. For example, PIP signaling at the plasma membrane requires the movement of phosphatidylinositol (PI) from the ER, where it is synthesized, to the plasma membrane to replenish PIP pools (Pemberton et al., 2020). Transport of PIPs out of the plasma membrane by LTPs, hypothetically, could also serve to attenuate PIP signaling. Nonvesicular lipid transport may also be required to enrich specific lipids in an organelle (Fig. 1 B). For example, nonvesicular transport of phosphatidylserine (PS) from the ER to the plasma membrane is required to enrich PS in the plasma membrane (Chung et al., 2015; Moser von Filseck et al., 2015a). Similarly, nonvesicular cholesterol transport from the ER to the Golgi complex is necessary to maintain cholesterol levels in the Golgi complex (de Saint-Jean et al., 2011; Mesmin et al., 2013). Both of these processes are driven by what has been termed counter-exchanging transport. This and other mechanisms of determining the directionality of lipid transport will be discussed later in this review. A third function for low-volume lipid transport is to regulate membrane organization at membrane contact sites (Fig. 1 B). In yeast, sterol-transporting LTPs at contact sites between the ER and vacuole—an organelle similar to lysosomes in higher eukaryotes—are necessary for the formation of sterol-enriched domains on the vacuole membrane during stress (Murley et al., 2017). More recently, it has been found that lipid trafficking by LTPs at contact sites between the ER and plasma membrane may promote the assembly of lipid nanodomains in the plasma membrane, which is critical for regulating vesicular trafficking to that compartment (Nishimura et al., 2019). It has been suggested that lipid microdomains are found at many contacts sites (King et al., 2020), and their formation may be driven by lipid transport by LTPs at these sites.While these and other studies show there is significant nonvesicular lipid transport at some MCSs, it is important to note that, for most MCSs, we still know little or nothing about how much lipid exchange occurs.Studying nonvesicular lipid exchange in cellsOur knowledge of nonvesicular lipid trafficking in cells is incomplete because following the movements of lipids in cells is challenging. Four types of approaches are used and each has notable limitations (Table 2). One is to use microscopy to follow the trafficking of fluorescent lipids in cells. These lipids, which are added exogenously to cells, contain fluorescent groups, like boron-dipyrromethene, for example (Marks et al., 2005), or are naturally fluorescent, like DHE (Hao et al., 2002). However, boron-dipyrromethene and other fluorescent groups can have biophysical properties that can make their intracellular trafficking and metabolism different from endogenous lipid—they can also perturb cells (Maekawa and Fairn, 2014). New tools for following lipid movements in cells are being rapidly developed that may overcome some of these limitations (Bumpus and Baskin, 2018). A second approach (Table 2) is to track lipids from their site of synthesis to other organelles. Radiolabeled or stable isotope–labeled lipid precursors are added to cells, and the transport of newly synthesized lipids is assessed by fractionating cells. This approach was first used in early studies of lipid transport to the plasma membrane (Kaplan and Simoni, 1985a; Kaplan and Simoni, 1985b); however, fractionation can be laborious and it is often difficult to obtain pure fractions. A third approach (Table 2) uses lipid modification in cells to indirectly measure the transport of newly synthesized lipids without the need for cellular fractionation. This approach requires localizing a lipid-modifying enzyme outside the organelle where lipid synthesis occurs. Lipid modification indicates that the newly synthesized lipid has been transported from its site of synthesis to the organelle containing the lipid-modifying enzyme. For example, PS transport from the ER to mitochondria has been estimated by measuring the conversion of newly synthesized PS, which is made in the ER, to phosphatidylethanolamine, which is catalyzed by an enzyme in mitochondria (Vance, 2015). There are a number of caveats to this type of approach: Lipid transport may not be the primary factor determining the rate of lipid modification, modified lipids may affect membrane function, and ensuring that the lipid-modifying enzyme is only, or mostly, active in the desired organelle can be challenging. A fourth approach (Table 2) is to use fluorescent lipid-binding proteins, often called lipid sensors, to measure changes in the lipid content of a membrane, which can occur as a result of lipid transport. These sensors contain a fluorescent protein fused to a protein domain that binds membranes containing a specific lipid (Wills et al., 2018). For example, some pleckstrin homology domains bind membrane containing specific PIP species. Pleckstrin homology domains fused to fluorescent proteins have been used to measure changes in PIP levels in cellular membranes (Várnai et al., 2017). There are important caveats to this approach. One is that the sensor can perturb membranes, and another is that the membrane affinity of some lipid sensors is determined by factors in addition to ligand concentration (Várnai et al., 2017). In addition, membrane binding by a sensor may be affected by changes in the availability of the lipid in a membrane, rather than changes in the lipid concentration of the membrane.Table 2.Approaches used to study lipid trafficking in cells
ApproachHow it worksAdvantagesCaveats
Fluorescent lipids-Fluorescent lipids added to cells -Trafficking assessed by microscopy-Monitor trafficking in real time in live cells-Fluorescent groups can alter physical properties and metabolism of lipids -Fluorescent lipids may affect cells
Labeling & fractionation-Labeled lipid precursors added to cells -Cells fractionated-Allows simultaneous analysis of transport and metabolism -Direct measurement of lipid levels in organelles-Fractionation challenging -Analysis requires cell destruction
Lipid modification-Lipid modifying enzyme localized in compartment outside synthesis site -Labeled lipid precursors added to cells-Does not require fractionation-Lipid modification must be faster than transport -Modified lipids may affect cells -Requires effective localization of modifying enzyme -Analysis requires cell destruction
Lipid sensorsFluorescent or luminescent sensors indicate changes in membrane lipid composition -Monitor trafficking in real time in live cells -Can detect small changes in membrane lipid composition-Factors other than lipid concentration may affect sensor binding -Sensors may perturb membranes
Open in a separate windowThe lipid transport machineryAs noted before, LTPs generally fall into one of two major categories, acting either as shuttles or as bridges to facilitate lipid movement through the cytosol between membranes (Fig. 2 A). Both types of lipid transporter feature a hydrophobic cavity that shields lipids as they transit the cytosol, though the cavity size differs. Shuttles mostly resemble cups in their overall shape, typically accommodating a single lipid moiety within the cup cavity, and often they have a lid that closes over the lipid once it has bound. Cup-like transporters first associate with the donor organelle to select and extract the cargo lipid and then ferry the lipid to and associate with the acceptor membrane, finally inserting the lipid there. In its open form, the lid may facilitate transporter association with organelle membranes, and hence lipid extraction from or deposition to the membrane, or—not mutually exclusive—the lid can play a role in the recognition of cargo lipids. The structure and function of these transporters, including the oxysterol-binding protein (OSBP)–related proteins (ORP/OSH), the StARkin, and some members of the tubular lipid binding proteins (TULIP) superfamily, have been excellently described (Wong et al., 2019). These cup-like transporters usually recognize either a single class of lipid or a limited subset of lipids, consistent with a role for these proteins in tweaking the membrane lipid compositions of different organelles (Fig. 2 B). The most efficient shuttles transfer approximately one phospholipid per second, at least in reconstituted systems (de Saint-Jean et al., 2011; Moser von Filseck et al., 2015a), too slow to yield the high-volume bulk lipid transport required for organelle biogenesis. It is possible that lipid-shuttling LTPs transport at higher rates in cells than they do in vitro. Factors that could boost or regulate LTP transport are rate discussed in the next section.Open in a separate windowFigure 2.Lipid transport machinery. (A) Schematic of a shuttling LTP (left) and a bridge-like LTP (right). A shuttling LTP (blue) extracts lipid monomers from one bilayer and then diffuses to a second bilayer and delivers the lipid. Bridge-like LTPs (light blue) form conduits that allow lipid molecules to flow between membranes. (B) Cup-like lipid transport modules with lipid bound. Osh4 can bind either sterol or PI4P (magenta) in the same pocket, with slight rearrangements in the lids (yellow). Osh4 is shown in different orientations. The StART-domain of CERT is shown with ceramide (magenta) bound. Protein Data Bank accession nos. are indicated: 1ZHZ, 3SPW, 2E3O, 6CBC, 6A9J, 5TV4, and 6MIT. (C) Bridge-like lipid transporters. From left to right: Intact VPS13 structure at ∼30-Å resolution by negative stain EM (De et al., 2017; courtesy of Y. Skiniotis, Stanford School of Medicine, Stanford, CA); cryo-EM structure of the N-terminal 160 kD of VPS13 showing it forms a tunnel (EMD-21113); ribbons representations of the VPS13 and ATG2 N-terminal fragments, showing they have the same fold; and ∼18-Å resolution cryo-EM structure of intact ATG2. (D) The LPS transport system in the inner membrane of E. coli, showing the flippase MsbA, which flips LPS from the inner to the outer leaflet of the membrane, and part of the transporter, which features an integral membrane portion that helps to load lipid into the bridge-like portion (indicated). ATPase domains in MsbA and in the LPS transporter are highlighted (light green and light blue).High-volume bulk lipid transfer could instead be mediated largely by bridge-like transporters, which may be able to transport lipids at significantly higher rates than cup-like LTPs. Previously known only in bacteria, these were recently also found in eukaryotes, and, so far, only a handful have been identified. Whether the endoplasmic reticulum–mitochondria encounter structure (ERMES) complex, which comprises several TULIP modules strung together into a tube and that mediates glycerolipid transport between the ER and mitochondria in yeast, is a shuttle or bridge-like transporter is under active discussion (Kornmann, 2020). Currently, the best studied of the bridge-like transporters are vacuolar protein sorting 13 homologue (Vps13) and autophagy-related protein 2 (Atg2; Chowdhury et al., 2018; De et al., 2017; Kumar et al., 2018; Li et al., 2020; Maeda et al., 2019; Osawa et al., 2019; Valverde et al., 2019), both of which are conserved across all eukaryotes. Most likely, they are evolutionarily related as they share a ∼120-residue chorein-N motif that forms part of the lipid transport module at their very N terminus (Kumar et al., 2018; Muñoz-Braceras et al., 2015; Osawa et al., 2019). They are large proteins long enough to span the 10–30-nm space between apposed membranes at membrane contact sites, allowing lipids to traverse the cytosolic space via a long hydrophobic groove that accommodates many lipids at once (Fig. 2 C). In contrast to the shuttles, these proteins may remain stably associated with both donor and acceptor membrane throughout the transfer process. Lipids are extracted from the donor membrane, then move along the bridge rather than occupying a well-defined binding site, as in cup-like proteins, and finally insert into the acceptor membrane. Vps13 resembles a bubble wand at low resolution, with a stem and a loop at one end, the latter probably corresponding to a predicted C-terminal WD40 (beta-transducin repeat) domain that plays a role in localization (De et al., 2017; Fig. 2 C). ATG2 is smaller, lacking the WD40 (beta-transducin repeat) domain and the corresponding “loop” (Chowdhury et al., 2018). A low-resolution cryo-EM reconstruction of ATG2 revealed a groove running along its entire length that could potentially serve as a track for lipids (Valverde et al., 2019; Fig. 2 C). A higher-resolution reconstruction for a ∼160-kD N-terminal fragment of Vps13 (∼40% of the protein) shows a series of β-strands assembled into a taco-like shell (Fig. 2 C) lined with hydrophobic residues along its entire length (Li et al., 2020). Both Vps13 and Atg2 bind to glycerolipids apparently nonspecifically (Kumar et al., 2018; Valverde et al., 2019), with the lipid fatty acyl moieties presumably bound within the taco and hydrophilic headgroups exposed in the solvent. In VPS13, the taco shell is widest at the N-terminal end and narrows toward the C-terminal end, where lipids most likely would be accommodated in a single file as they flow through the protein. Among other roles, both Vps13 and Atg2 play roles in membrane expansion. They are critical for the biogenesis of at least three double-membrane, cup-shaped structures: the yeast pro-spore membrane during sporulation, the acrosome that forms at the tip of spermatids (Da Costa et al., 2020), and the autophagosome (Park and Neiman, 2012; Suzuki et al., 2013b); Vps13A in humans is proposed to play a role in ER–mitochondrial glycerolipid exchange (Kumar et al., 2018).While we argue here that bridge-like LTPs may transport significantly faster than shuttling transports in cells, it is important to note that, to date, this has not been found in vitro (Kumar et al., 2018; Li et al., 2020; Osawa et al., 2019; Valverde et al., 2019). We consider it likely that these proteins do not function in isolation, but rather as part of lipid transport systems, as considered below. Still unidentified, the additional components have not been included in assays in vitro. In addition, one study (Valverde et al., 2019) shows that an N-terminal fragment of the bridge-forming LTP Atg2 can rescue the full-length protein when 10-fold overexpressed in cells. This fragment likely can function as a shuttle, probably because the Atg2 N terminus is also responsible for Atg2 localization (Tamura et al., 2017). How both shuttling and bridge-forming LTPs work in conjunction with other proteins in cells remains to be determined.Studies of eukaryotic lipid transfer have focused primarily on the ability of LTPs to move lipids between membranes, either working alone or together with integral membrane proteins, like the vesicle-associated membrane protein–associated proteins (VAPs), which are ER-resident proteins. There has been little investigation of roles for integral membrane proteins or domains as active participants in lipid transfer rather than simply as scaffolds to localize LTPs. In contrast, LTPs responsible for lipid transport across the periplasm in gram-negative bacteria frequently function as components of systems, including several multispan integral membrane components, which facilitate lipid extraction from or insertion into membranes. For example, lipopolysaccharide (LPS) transport across the periplasm involves a β-strand bridge—comprising the proteins LptC, LptA, and soluble portions of LptD—reminiscent of VPS13, with a groove lined with hydrophobics (Li et al., 2019; Owens et al., 2019; Fig. 2 D). Additionally, the LPS transport system includes an ATP-binding cassette transporter embedded in the inner membrane that selectively loads LPS onto the bridge, ensuring directional flow toward the outer membrane, and, further, an integral membrane protein in the outer membrane, LptD, which facilitates lipid transfer from the inner to the outer leaflet of this membrane (Sperandeo et al., 2017). Collaborations between a bridge-like transporter or a shuttle, respectively, and integral membrane proteins are also observed in the case of the Escherichia coli Pql and Mla phospholipid transport systems (Ekiert et al., 2017; Isom et al., 2020).Similarly, lipid transfer in eukaryotes by bridge-like transporters and some shuttling transporters might also involve integral membrane partners. Indeed, multispan integral membrane proteins were recently discovered to play a role in some instances of cholesterol transfer. Cholesterol transfer from the lumen of lysosomes to the lysosomal membrane has been particularly well characterized and requires Niemann-Pick proteins, NPC1 and NPC2. NPC2 is a cup-like LTP shuttle that solubilizes sterols in the lysosomal lumen and hands them off to the luminal domain of NPC1, which also has a membrane-embedded portion. The luminal domain of NPC1 forms a hydrophobic channel that funnels cholesterol through the lysosomal glycocalyx, a carbohydrate-enriched coating that covers the inside surface of lysosomes. The cholesterol is then handed off to the multipass membrane domain of NPC1, which facilitates membrane insertion (Gong et al., 2016; Li et al., 2016; Winkler et al., 2019). A beautiful recent study revealed mechanistic details of cholesterol transfer from NPC2 to NPC1 and demonstrated that it is pH dependent, which is consistent with the low pH in the lumen of lysosomes, where it occurs (Qian et al., 2020). A similar arrangement has also been observed in hedgehog signaling, except that here membrane-embedded portions of the Patched receptor are thought to be involved in cholesterol extraction from the plasma membrane rather than sterol insertion (Gong et al., 2018; Qi et al., 2018; Zhang et al., 2018). Whether and how significantly the membrane-embedded protein domains accelerate cholesterol transfer remains unknown.Do eukaryotic glycerolipid transporters, like their bacterial counterparts, work together with partner proteins in the membrane that do more than simply localize them to contact sites? Direct interactions have been reported between both Vps13 and Atg2 and multispan integral membrane proteins (Guardia et al., 2020; John Peter et al., 2017; Tang et al., 2019), and others may yet be identified. Interestingly, the interaction of Vps13 with the mitochondria outer membrane protein, Mcp1, seems to be required for more than Vps13 localization, though how Mcp1 affects Vps13 function remains to be determined (John Peter et al., 2017). The ERMES complex, which is involved in glycerolipid exchange between the ER and mitochondria in yeast, includes a multispan integral membrane protein (Mdm10; Kornmann et al., 2009). These membrane residents might have roles in selecting and/or extracting lipids from membranes, in loading lipids onto LTPs to drive directional transfer, in lipid insertion at the acceptor membrane, or in lipid scrambling or flipping (lipid movement between the two leaflets of a bilayer). Interestingly, there is growing evidence that some LTPs are associated with scramblases, integral membrane proteins that facilitate lipid movement between the leaflets of membrane bilayers. The bridge-forming Atg2 interacts with Atg9 (Guardia et al., 2020), which has recently been shown to be a scramblase and is key to autophagosome membrane expansion (Maeda et al., 2020; Matoba et al., 2020; Orii et al., 2021). The yeast cup-like phosphatidylserine transporter, Osh6, has been reported to work in conjunction with Ist2 (D’Ambrosio et al., 2020), which is homologous to members of the transmembrane protein 16 (TMEM16) family of Ca2+-activated scramblases (Suzuki et al., 2013a). Conceivably, membrane proteins, like scramblases, that associate with LTPs could play a role in accelerating lipid transfer beyond the rates observed for LTPs operating by themselves in vitro.How do LTPs facilitate transport?How LTPs facilitate lipid transport between membranes has been studied mostly in the context of cup-like transporters, though many of the same principles may also apply to bridge-like transporters. The basic outline of how these cup-like LTPs function, described in the previous section (Fig. 2 A), has been known for some time, but the mechanistic details are only now beginning to emerge. To understand how LTPs extract lipids from membranes, it is important to consider the energetics of lipid desorption from membranes. Lipids can desorb from membranes spontaneously without the assistance of proteins, but the rates are very slow. For example, the half-time of the exchange of phospholipids between liposomes is tens of hours (Jones and Thompson, 1990; McLean and Phillips, 1981; McLean and Phillips, 1984). This is because there is a high energic barrier to spontaneous lipid movement into the aqueous phase. One study estimated that the desorption free energy of pulling a phospholipid entirely out of a bilayer is ∼63 kJ⋅mol−1 (Grafmüller et al., 2013). LTPs facilitate lipid transfer by significantly lowering the energy of lipid desorption from membrane, because the lipid is no longer desorbed into the aqueous phase, but rather into the hydrophobic cavity of the protein.The molecular mechanism by which LTPs remove and deliver lipid monomers to membranes remain to be determined; both probably require LTPs to partially insert into membranes or disrupt bilayer organization. Another important question is how the affinity of LTPs for membranes and lipids affects the rate of lipid extraction and delivery to membranes. Since many LTPs contain flexible lid-like domains that shield bound lipids, the opening and closing of these lids may play critical roles in determining lipid extraction and delivery. For example, the lid of the LTP Osh4/Kes1 regulates lipid transport, preventing the release of one type of cargo but not of a second (Moser von Filseck et al., 2015b).Given how challenging it is to understand the interactions of LTPs with membranes at a molecular level, molecular dynamics simulations may currently be the best hope of gaining insight. Simulations of membrane binding and phospholipid extraction by two cup-like LTPs revealed that both have domains that penetrate the bilayer upon membrane binding and may help orient the proteins to facilitate lipid extraction (Grabon et al., 2017; Miliara et al., 2019). Unfortunately, in both cases, the phospholipids were not fully extracted during the simulations, which were run for up to 5 µs, and the details of how these LTPs remove phospholipids from membranes remain to be elucidated. LTPs probably capture lipids during what have been termed protrusions (Pfeiffer, 2015), spontaneously occurring excursions of lipid molecules partially out of a membrane (i.e., partial desorption). LTPs may facilitate lipid protrusions from membranes by penetrating or binding the bilayer.Since the propensity of lipids to protrude from membranes is determined by the biophysical properties of bilayers, such as hydration, curvature, tension, lipid packing, and order, these factors must also affect lipid transport by LTPs. Proteins and lipids at contact sites that determine these properties could have a significant effect on lipid transport. We are just beginning to understand the physical properties of membranes at MCSs. Two recent studies showed that, in yeast, the ER at some ER–plasma membrane contact sites forms peaks with extremely high curvature that could affect lipid exchange at these sites; peak formation requires tricalbins, the yeast orthologues of extended synaptotagmins in mammals (Collado et al., 2019; Hoffmann et al., 2019). Other ER-shaping proteins, the reticulons, have also been proposed to promote lipid exchange at ER–mitochondria contacts (Voss et al., 2012). There are also hints that the lipid composition of the ER at MCSs can differ from the rest of an ER, which could affect lipid transport at these sites. It has been suggested that ER–mitochondria contacts and perhaps other contact sites have raft-like properties (Area-Gomez et al., 2012; Currinn et al., 2016; King et al., 2020). If the membranes at these MCSs are, like rafts, more ordered than the surrounding membrane, this would be expected to inhibit lipid protrusion and reduced transport. There is also some indication that lipid production at contact sites facilitates transport (Kannan et al., 2017; Schütter et al., 2020), perhaps by increasing the frequency of lipid protrusions at these sites. Understanding the biophysical properties of membranes at contact sites and how they modulate lipid exchange remains an important challenge for the future.How LTPs deliver lipids to membranes is another important question. It is likely that the same factors that affect lipid partitioning from donor membrane into the LTP also affect lipid transfer from the LTP into the receiving membrane. A recent molecular dynamics simulation of sterol exit from a StARkin LTP suggests that it is facilitated by the entry of water into the lipid-binding pocket (Khelashvili et al., 2019). This may be true of other sterol-binding LTPs (Singh et al., 2009), but it remains to be determined whether this mechanism holds for other classes of lipids and types of LTPs.What are the mechanisms underlying directional transport?Cup-like LTPs do not require energy to transport lipids between liposomes and can spontaneously equilibrate their cargos between two populations of liposomes; however, in cells, some LTP-mediated lipid transport is directional. A number of mechanisms are employed (Fig. 3). In some cases, directional transfer occurs when a lipid is moved from one membrane to a second, but cannot be returned, either because it is enzymatically altered (Fig. 3 A) or because it becomes complexed with other lipids in the second membrane (Fig. 3 B). For example, the ceramide transfer protein (CERT) shuttles ceramide from the ER to the Golgi complex, where the ceramide is converted to complex sphinoglipids, which CERT cannot transport back to the ER (Hanada et al., 2003). Directional lipid transport by LTPs may also be promoted by lipid synthesis in one of the membranes at an MCS (Fig. 3 C). Phospholipid synthesis in regions of the ER in contact with mitochondria or growing phagophores may promote transport to these organelles (Kannan et al., 2017; Schütter et al., 2020). Directional transport can also be driven by ATP hydrolysis. In E. coli, the ATP-binding cassette transporter LptBFG may use ATP hydrolysis to drive LPS out of the inner membrane into a bridge-like domain, and then into the outer membrane (Owens et al., 2019; Figs. 2 D and and3).3). In other words, the transporter may push lipids into a tube-like domain, which causes directional transport. It is possible that other transporters similarly drive directional transport by using ATP hydrolysis to pull lipids out of a tube-like domain and into a membrane. One such transporter may be the trigalactosyldiacylglycerol in Arabidopsis thaliana, which spans the inner and outer chloroplast membrane. This complex is thought to use ATP hydrolysis to drive lipids phospholipids from the ER into chloroplasts (Fan et al., 2015; Roston et al., 2012; Wang et al., 2012). Whether the trigalactosyldiacylglycerol complex forms a tube-like lipid transporter and the role of ATP hydrolysis in driving transport remains to be determined.Open in a separate windowFigure 3.Examples of five mechanisms of directional transport. (A) Lipid consumption–driven lipid transport. Ceramide is transported by CERT from the ER to the Golgi, where it is converted to sphingomyelin (SM) and other complex sphingolipids and cannot be returned to the ER by CERT. (B) Lipid domain formation can drive lipid transport when a lipid becomes associated with a domain in one of the two membranes. Lam6/Ltc1 brings the sterol, ergosterol, to the vacuole in S. cerevisiae, and membrane domain formation probably drives the accumulation of ergosterol (domain in red). (C) Lipid synthesis at MCSs can drive lipid transport. PS synthesis at ER–mitochondria contact sites in S. cerevisiae promotes PS transport to mitochondria. (D) ATP consumption drives LPS transport from the inner to the outer membrane of E. coli. (E) Counter-exchange transport using the difference in PIP concentration in two membranes to drive the transport of a second lipid. OSBP uses counter-exchange transport to enrich cholesterol in the Golgi membrane.Some of the transporters in the ORP/Osh family make use of energy stored in cellular phosphoinositide gradients to redistribute lipids from their sites of synthesis in the ER to other organelles, where these lipids are enriched. One of the best characterized examples is OSBP, which transfers sterol to the Golgi, where cholesterol levels are higher relative to the ER, by counter-exchanging ER-derived sterol for phosphatidylinositol 4-phosphate (PI4P) present at the Golgi (Mesmin et al., 2013). OSBP exchanges its sterol cargo for PI4P at the Golgi because the LTP has a higher affinity for the phosphoinositide. The LTP then carries the phosphoinositide back to the ER, where a resident lipid phosphatase (Sac1) hydrolyzes it to PI to decrease its affinity for the LTP and to allow the LTP to pick up another sterol molecule (de Saint-Jean et al., 2011; Mesmin et al., 2013). Phosphatidylserine transfer from the ER to the plasma membrane is mediated by Orp5/Orp8 (Osh6 in yeast) by using a similar mechanism, exchanging the glycerolipid for PI4P or phosphatidylinositol 4,5-bisphosphate (PI(4,5)P2) present at the plasma membrane (Chung et al., 2015; Ghai et al., 2017; Moser von Filseck et al., 2015a). While it is possible that ORP/OSH proteins are the only family of transporters that operate in this way, the mechanism could be more general, extending to other LTP families as well, and there may be additional, still undiscovered mechanisms by which cup-like LTPs to transport lipids against their gradient. While counterexchange is widely viewed as a mechanism to redistribute ER-derived lipids against a gradient, it may also serve to regulate phosphoinositide levels at the acceptor membrane and may play a role in attenuating or regulating signaling.The most extreme case of directional transfer occurs during organelle biogenesis, likely facilitated by bridge-like LTPs, such as VPS13 or ATG2. As these LTPs seem to bind glycerolipids nonspecifically and, moreover, simultaneously bind tens of these lipids that must all move concertedly for efficient transport (Kumar et al., 2018; Valverde et al., 2019), it seems unlikely that chemical gradients could drive directionality. Instead, lipids are probably pushed or pulled into bridge-like transporters, perhaps driven by lipid synthesis or ATP hydrolysis. Since LTPs transfer lipids between the cytosolic leaflets of different organelles, they almost certainly operate together with scramblases or flippases that can transfer lipids between the cytosolic and luminal leaflets, since membrane expansion requires both leaflets to grow. It is not known whether these scramblases/flippases must be associated directly with the LTP, as in the LPS system. Discovering the basis of directional transfer by bridge-like proteins, potentially by identifying interacting partners in this process, is an ambitious undertaking for the coming years.Conclusions and future directionsThe last 10 yr have seen tremendous progress in our understanding of nonvesicular lipid transport in cells in both its functions and mechanisms; however, many fundamental questions remain. Our knowledge of the volume and rates of nonvesicular lipid transport at MCSs is still incomplete, primarily because measuring lipid movements in cells remains challenging. A better picture is likely to emerge in the next few years as new lipid sensors and fluorescent lipid analogues are developed and used with rapidly evolving super-resolution microscopy.A better knowledge of how LTPs promote lipid desorption and delivery to membranes will be critical for understanding how lipid transfer rates in cells are determined. The mechanistic and energetic details are not well understood for any LTP, let alone representatives from all of the various LTP families. Molecular dynamics simulations may lead the way. Another important question is how lipids are moved within tube-like transporters or between LTPs and other proteins. LTPs have largely been studied in isolation, but in cells, their activity is probably determined by proteins that affect the properties of membranes, such as those that deform membranes—lipid metabolizing enzymes and flippases (or scramblases). These proteins may also help to determine the directionality of lipid transport by LTPs.Understanding how high-volume bulk lipid transport occurs is another important challenge for the field. It seems likely that LTPs transport lipids significantly faster in cells than they do in vitro, but how is not clear. We still have much to learn about the microenvironments formed at MCSs and how they contribute to lipid exchange by LTPs. In the next few years, we are likely to get a better understanding of how LTPs function and work in concert with other proteins that modulate membranes.  相似文献   

19.
20.
The evolution of Na+-selective four-domain voltage-gated channels (4D-Navs) in animals allowed rapid Na+-dependent electrical excitability, and enabled the development of sophisticated systems for rapid and long-range signaling. While bacteria encode single-domain Na+-selective voltage-gated channels (BacNav), they typically exhibit much slower kinetics than 4D-Navs, and are not thought to have crossed the prokaryote–eukaryote boundary. As such, the capacity for rapid Na+-selective signaling is considered to be confined to certain animal taxa, and absent from photosynthetic eukaryotes. Certainly, in land plants, such as the Venus flytrap (Dionaea muscipula) where fast electrical excitability has been described, this is most likely based on fast anion channels. Here, we report a unique class of eukaryotic Na+-selective, single-domain channels (EukCatBs) that are present primarily in haptophyte algae, including the ecologically important calcifying coccolithophores, Emiliania huxleyi and Scyphosphaera apsteinii. The EukCatB channels exhibit very rapid voltage-dependent activation and inactivation kinetics, and isoform-specific sensitivity to the highly selective 4D-Nav blocker tetrodotoxin. The results demonstrate that the capacity for rapid Na+-based signaling in eukaryotes is not restricted to animals or to the presence of 4D-Navs. The EukCatB channels therefore represent an independent evolution of fast Na+-based electrical signaling in eukaryotes that likely contribute to sophisticated cellular control mechanisms operating on very short time scales in unicellular algae.

Electrical signals trigger rapid physiological events that underpin an array of fundamental processes in eukaryotes, from contractile amoeboid locomotion (Bingley and Thompson, 1962), to the action potentials of mammalian nerve and muscle cells (Hodgkin and Huxley, 1952). These events are mediated by voltage-gated ion channels (Brunet and Arendt, 2015). In excitable animal cells, Ca2+- or Na+-selective members of the four-domain voltage-gated cation channel family (4D-Cav/Nav) underpin well-characterized signaling processes (Catterall et al., 2017). The 4D-Cav/Nav family is broadly distributed across eukaryotes, contributing to signaling processes associated with motility in some unicellular protist and microalgal species (Fujiu et al., 2009; Lodh et al., 2016), although these channels are absent from land plants (Edel et al., 2017). It is likely that the ancestral 4D-Cav/Nav channel was Ca2+-permeable, with Na+-selective channels arising later within the animal lineage (Moran et al., 2015). This shift in ion selectivity represented an important innovation in animals, allowing rapid voltage-driven electrical excitability to be decoupled from intracellular Ca2+ signaling processes (Moran et al., 2015).Na+-selective voltage-gated channels have not been described in other eukaryotes, although a large family of Na+-selective channels (BacNav) is present in prokaryotes (Ren et al., 2001; Koishi et al., 2004). BacNav are single-domain channels that form homotetramers, resembling the four-domain architecture of 4D-Cav/Nav. Studies of BacNav channels have provided considerable insight into the mechanisms of gating and selectivity in voltage-dependent ion channels (Payandeh et al., 2012; Zhang et al., 2012). The range of activation and inactivation kinetics of native BacNav are generally slower than observed for 4D-Nav, suggesting that the concatenation and subsequent differentiation of individual pore-forming subunits may have enabled 4D-Nav to develop specific properties such as fast inactivation, which is mediated by the conserved intracellular Ile–Phe–Met linker between domains III and IV (Fig. 1A; Irie et al., 2010; Catterall et al., 2017).Open in a separate windowFigure 1.EukCatBs represent a novel class of single-domain channels. A, Schematic diagram of a single-domain EukCatB channel. The voltage-sensing module (S1–S4, blue), including conserved positively charged (++) residues of segment (S4) that responds to changes in membrane potential, is shown. The pore module (S5–S6, red) is also indicated, including the SF motif (Ren et al., 2001). The structure of a 4D-Nav (showing the SF of rat 4D-Nav1.4 with canonical “DEKA” locus of Na+-selective 4D-Nav1s) is also displayed (right). The Ile–Phe–Met motif of the fast inactivation gate is indicated (West et al., 1992) B, Maximum likelihood phylogenetic tree of single-domain, voltage-gated channels including BacNav and the three distinct classes of EukCat channels (EukCatA–C). Representatives of the specialized family of single-domain Ca2+ channels identified in mammalian sperm (CatSpers) are also included. SF for each sequence is shown (right). “Position 0” of the high-field–strength site that is known to be important in determining Na+ selectivity (Payandeh et al., 2011), is colored red. Channel sequences selected for functional characterization in this study are shown in bold. EukCatA sequences previously characterized (Helliwell et al., 2019) are also indicated, as is NaChBac channel from B. halodurans (Ren et al., 2001). Maximum likelihood bootstrap values (>70) and Bayesian posterior probabilities (>0.95) are above and below nodes, respectively. Scanning electron micrographs of coccolithophores E. huxleyi (scale bar = 2 μm) and S. apsteinii, (scale bar = 10 μm) are shown.We recently identified several classes of ion channel (EukCats) bearing similarity to BacNav in the genomes of eukaryotic phytoplankton. Characterization of EukCatAs found in marine diatoms demonstrated that these voltage-gated channels are nonselective (exhibiting permeability to both Na+ and Ca2+) and play a role in depolarization-activated Ca2+ signaling (Helliwell et al., 2019). Two other distinct classes of single-domain channels (EukCatBs and EukCatCs) were also identified that remain uncharacterized. These channels are present in haptophytes, pelagophytes, and cryptophytes (EukCatBs), as well as dinoflagellates (EukCatCs; Helliwell et al., 2019). Although there is a degree of sequence similarity between the distinct EukCat clades, the relationships between clades are not well resolved, and there is not clear support for a monophyletic origin of EukCats. The diverse classes of EukCats may therefore exhibit significant functional differences. Characterization of these different classes of eukaryote single-domain channels is thus vital to our understanding of eukaryote ion channel structure, function, and evolution, and to our gaining insight into eukaryote membrane physiology more broadly.Notably, EukCatB channels were found in ecologically important coccolithophores, a group of unicellular haptophyte algae that represent major primary producers in marine ecosystems. Coccolithophores are characterized by their ability to produce a cell covering of ornate calcium carbonate platelets (coccoliths; Fig. 1B; Taylor et al., 2017). The calcification process plays an important role in global carbon cycling, with the sinking of coccoliths representing a major flux of carbon to the deep ocean. Patch-clamp studies of coccolithophores indicate several unusual aspects of membrane physiology, such as an inwardly rectifying Cl conductance and a large outward H+ conductance at positive membrane potentials, which may relate to the increased requirement for pH homeostasis associated with intracellular calcification. Here we report that EukCatB channels from two coccolithophore species (Emiliania huxleyi and Scyphosphaera apsteinii) act as very fast Na+-selective voltage-gated channels that exhibit many similarities to the 4D-Navs, which underpin neuronal signaling in animals. Thus, our findings demonstrate that the capacity for rapid Na+-based signaling has evolved in certain photosynthetic eukaryotes, contrary to previous widely held thinking.  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