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 109 毫秒
1.
2.
3.
4.
5.
6.
7.
8.
9.
Time-lapse fluorescence microscopy is one of the main tools used to image subcellular structures in living cells. Yet for decades it has been applied primarily to in vitro model systems. Thanks to the most recent advancements in intravital microscopy, this approach has finally been extended to live rodents. This represents a major breakthrough that will provide unprecedented new opportunities to study mammalian cell biology in vivo and has already provided new insight in the fields of neurobiology, immunology, and cancer biology.The discovery of GFP combined with the ability to engineer its expression in living cells has revolutionized mammalian cell biology (Chalfie et al., 1994). Since its introduction, several light microscopy–based techniques have become invaluable tools to investigate intracellular events (Lippincott-Schwartz, 2011). Among them are: time-lapse confocal microscopy, which has been instrumental in studying the dynamics of cellular and subcellular processes (Hirschberg et al., 1998; Jakobs, 2006; Cardarelli and Gratton, 2010); FRAP, which has enabled determining various biophysical properties of proteins in living cells (Berkovich et al., 2011); and fluorescence resonance energy transfer (FRET), which has been used to probe for protein–protein interactions and the local activation of specific signaling pathways (Balla, 2009). The continuous search for improvements in temporal and spatial resolution has led to the development of more sophisticated technologies, such as spinning disk microscopy, which allows the resolution of fast cellular events that occur on the order of milliseconds (Nakano, 2002); total internal reflection microscopy (TIRF), which enables imaging events in close proximity (100 nm) to the plasma membrane (Cocucci et al., 2012); and super-resolution microscopy (SIM, PALM, and STORM), which captures images with resolution higher than the diffraction limit of light (Lippincott-Schwartz, 2011).Most of these techniques have been primarily applied to in vitro model systems, such as cells grown on solid substrates or in 3D matrices, explanted embryos, and organ cultures. These systems, which are relatively easy to maintain and to manipulate either pharmacologically or genetically, have been instrumental in providing fundamental information about cellular events down to the molecular level. However, they often fail to reconstitute the complex architecture and physiology of multicellular tissues in vivo. Indeed, in a live organism, cells exhibit a 3D organization, interact with different cell types, and are constantly exposed to a multitude of signals originated from the vasculature, the central nervous system, and the extracellular environment. For this reason, scientists have been attracted by the possibility of imaging biological processes in live multicellular organisms (i.e., intravital microscopy [IVM]). The first attempt in this direction was in 1839, when Rudolph Wagner described the interaction of leukocytes with the walls of blood vessels in the webbed feet of a live frog by using bright-field transillumination (Wagner, 1839). Since then, this approach has been used for over a century to study vascular biology in thin areas of surgically exposed organs (Irwin and MacDonald, 1953; Zweifach, 1954) or by implanting optical windows in the skin or the ears (Clark and Clark, 1932). In addition, cell migration has also been investigated using transparent tissues, such as the fin of the teleost (Wood and Thorogood, 1984; Thorogood and Wood, 1987). The introduction of epifluorescence microscopy has enabled following in more detail the dynamics of individual cells in circulation (Nuttall, 1987), in tumors (MacDonald et al., 1992), or in the immune system (von Andrian, 1996), and the spatial resolution has been significantly improved by the use of confocal microscopy, which has made it possible to collect serial optical sections from a given specimen (Villringer et al., 1989; O’Rourke and Fraser, 1990; Jester et al., 1991). However, these techniques can resolve structures only within a few micrometers from the surface of optically opaque tissues (Masedunskas et al., 2012a). It was only in the early nineteen nineties, with the development of multiphoton microscopy, that deep tissue imaging has become possible (Denk et al., 1990; Zipfel et al., 2003b), significantly contributing to several fields, including neurobiology, immunology, and cancer biology (Fig. 1; Svoboda and Yasuda, 2006; Amornphimoltham et al., 2011; Beerling et al., 2011). In the last few years, the development of strategies to minimize the motion artifacts caused by the heartbeat and respiration has made it possible to successfully image subcellular structures with spatial and temporal resolutions comparable to those achieved in in vitro model systems, thus providing the opportunity to study cell biology in live mammalian tissues (Fig. 1; Weigert et al., 2010; Pittet and Weissleder, 2011).Open in a separate windowFigure 1.Spatial resolution and current applications of intravital microscopy. IVM provides the opportunity to image several biological processes in live animals at different levels of resolution. Low-magnification objectives (5–10×) enable visualizing tissues and their components under physiological conditions and measuring their response under pathological conditions. Particularly, the dynamics of the vasculature have been one of topic most extensively studied by IVM. Objectives with higher magnification (20–30×) have enabled imaging the behavior of individual cell over long periods of time. This has led to major breakthroughs in fields such as neurobiology, immunology, cancer biology, and stem cell research. Finally, the recent developments of strategies to minimize the motion artifacts caused by the heartbeat and respiration combined with high power lenses (60–100×) have opened the door to image subcellular structures and to study cell biology in live animals.The aim of this review is to highlight the power of IVM in addressing cell biological questions that cannot be otherwise answered in vitro, due to the intrinsic limitations of reductionist models, or by other more classical approaches. Furthermore, we discuss limitations and areas for improvement of this imaging technique, hoping to provide cell biologists with the basis to assess whether IVM is the appropriate choice to address their scientific questions.

Imaging techniques currently used to perform intravital microscopy

Confocal and two-photon microscopy are the most widely used techniques to perform IVM. Confocal microscopy, which is based on single photon excitation, is a well-established technique (Fig. 2 A) that has been extensively discussed elsewhere (Wilson, 2002); hence we will only briefly describe some of the main features of two-photon microscopy and other nonlinear optical techniques.Open in a separate windowFigure 2.Fluorescent light microscopy imaging techniques used for intravital microscopy. (A) Confocal microscopy. (top) In confocal microscopy, a fluorophore absorbs a single photon with a wavelength in the UV-visible range of the spectrum (blue arrow). After a vibrational relaxation (orange curved arrow), a photon with a wavelength shifted toward the red is emitted (green arrow). (center) In thick tissue, excitation and emission occur in a relative large volume around the focal plane (F.P.). The off-focus emissions are eliminated through a pinhole, and the signal from the focal plane is detected via a photomultiplier (PMT). Confocal microscopy enables imaging at a maximal depth to 80–100 µm. (bottom) Confocal z stack of the tongue of a mouse expressing the membrane marker m-GFP (green) in the K14-positive basal epithelial layer, and the membrane marker mTomato in the endothelium (red). The xy view shows a maximal projection of 40 z slices acquired every 2.5 µm, whereas the xz view shows a lateral view of the stack. In blue are the nuclei labeled by a systemic injection of Hoechst. Excitation wavelengths: 450 nm, 488 nm, and 562 nm. (B) Two- and three-photon microscopy. (top) In this process a fluorophore absorbs almost simultaneously two or three photons that have half (red arrow) or a third (dark red arrow) of the energy required for its excitation with a single photon. Two- or three-photon excitations typically require near-IR or IR light (from 690 to 1,600 nm). (center) Emission and excitation occur only at the focal plane in a restricted volume (1.5 fl), and for this reason a pinhole is not required. Two- and three-photon microscopy enable imaging routinely at a maximal depth of 300–500 µm. (bottom) Two-photon z stack of an area adjacent to that imaged in A. xy view shows a maximal projection of 70 slices acquired every 5 µm. xz view shows a lateral view of the stack. Excitation wavelength: 840 nm. (C) SHG and THG. (top) In SHG and THG, photons interact with the specimen and combine to form new photons that are emitted with twice or three times their initial energy without any energy loss. (center) These processes have similar features to those described for two- and three-photon microscopy and enable imaging at a maximal depth of 200–400 µm. (bottom) z stack of a rat heart excited by two-photon microscopy (740 nm) to reveal the parenchyma (green), and SHG (930 nm) to reveal collagen fibers (red). xy shows a maximal projection of 20 slices acquired every 5 µm. xz view shows a lateral view of the stack. Bars: (xy views) 40 µm; (xz views) 50 µm.The first two-photon microscope (Denk et al., 1990) was based on the principle of two-photon excitation postulated by Maria Göppert-Mayer in her PhD thesis (Göppert-Mayer, 1931). In this process a fluorophore is excited by the simultaneous absorption of two photons with wavelengths in the near-infrared (IR) or IR spectrum (from 690 to 1,600 nm; Fig. 2 B). Two-photon excitation requires high-intensity light that is provided by lasers generating very short pulses (in the femtosecond range) and is focused on the excitation spot by high numerical aperture lenses (Zipfel et al., 2003b). There are three main advantages in using two-photon excitation for IVM. First, IR light has a deeper tissue penetration than UV or visible light (Theer and Denk, 2006). Indeed, two-photon microscopy can resolve structures up to a depth of 300–500 µm in most of the tissues (Fig. 2 B), and up to 1.5 mm in the brain (Theer et al., 2003; Masedunskas et al., 2012a), whereas confocal microscopy is limited to 80–100 µm (Fig. 2 A). Second, the excitation is restricted to a very small volume (1.5 fl; Fig. 2 B). This implies that in two-photon microscopy there is no need to eliminate off-focus signals, and that under the appropriate conditions photobleaching and phototoxicity are negligible (Zipfel et al., 2003b). However, confocal microscopy induces out-of-focus photodamage, and thus is less suited for long-term imaging. Third, selected endogenous molecules can be excited, thus providing the contrast to visualize specific biological structures without the need for exogenous labeling (Zipfel et al., 2003a). Some of these molecules can also be excited by confocal microscopy using UV light, although with the risk of inducing photodamage.More recently, other nonlinear optical techniques have been used for IVM, and among them are three-photon excitation, and second and third harmonic generation (SHG and THG; Campagnola and Loew, 2003; Zipfel et al., 2003b; Oheim et al., 2006). Three-photon excitation follows the same principle as two-photon (Fig. 2 B), and can reveal endogenous molecules such as serotonin and melatonin (Zipfel et al., 2003a; Ritsma et al., 2013). In SHG and THG, photons interact with the specimen and combine to form new photons that are emitted with two or three times their initial energy (Fig. 2 C). SHG reveals collagen (Fig. 2 C) and myosin fibers (Campagnola and Loew, 2003), whereas THG reveals lipid droplets and myelin fibers (Débarre et al., 2006; Weigelin et al., 2012). Recently, two other techniques have been used for IVM: coherent anti-Stokes Raman scattering (CARS) and fluorescence lifetime imaging (FLIM). CARS that is based on two laser beams combined to match the energy gap between two vibrational levels of the molecule of interest, has been used to image lipids and myelin fibers (Müller and Zumbusch, 2007; Fu et al., 2008; Le et al., 2010). FLIM, which measures the lifetime that a molecule spends in the excited state, provides quantitative information on cellular parameters such as pH, oxygen levels, ion concentration, and the metabolic state of various biomolecules (Levitt et al., 2009; Provenzano et al., 2009; Bakker et al., 2012).We want to emphasize that two-photon microscopy and the other nonlinear techniques are the obligatory choice when the imaging area is located deep inside the tissue, endogenous molecules have to be imaged, or long-term imaging with frequent sampling is required. However, confocal microscopy is more suited to resolve structures in the micrometer range, because of the possibility of modulating the optical slice (Masedunskas et al., 2012a).

IVM to investigate biological processes at the tissue and the single cell level

The main strength of IVM is to provide information on the dynamics of biological processes that otherwise cannot be reconstituted in vitro or ex vivo. Indeed, IVM has been instrumental in studying several aspects of tissue physiopathology (Fig. 3, A and B). Although other approaches such as classical immunohistochemistry, electron microscopy, and indirect immunofluorescence may provide detailed structural and quantitative information on blood vessels, IVM enables measuring events such as variations of blood flow at the level of the capillaries or local changes in blood vessel permeability. These data have been instrumental in understanding the mechanisms of ischemic diseases and tumor progression, and in designing effective anticancer treatments.

Table 1.

IVM to study tissue physiopathology
EventOrganProbesReference
Measurements of local blood flow and glial cell functionBrainDextranHelmchen and Kleinfeld, 2008
Ischemia and reperfusionBrainSulphorhodamine 101, DextranZhang and Murphy, 2007; Masamoto et al., 2012;
Glomerular filtration and tubular reabsorptionKidneyDextran, AlbuminKang et al., 2006; Yu et al., 2007; Camirand et al., 2011
Blood flow patternsPancreatic isletsDextranNyman et al., 2008
Capillary response and synaptic activationOlfactory bulbDextranChaigneau et al., 2003
Imaging angiogenesis during wound healingSkullcapDextranHolstein et al., 2011
Pulmonary microvasculature and endothelial activationLungDextranPresson et al., 2011
Morphology of blood vessels and permeability in tumorsXenograftsDextran, RGD quantum dotsTozer et al., 2005; Smith et al., 2008; Vakoc et al., 2009; Fukumura et al., 2010
Hepatic transport into the bile canaliculiLiverCarboxyfluorescein diacetate Rhodamine 123Babbey et al., 2012; Liu et al., 2012
Progression of amyloid plaques in Alzheimer’s diseaseBrainCurcumin and metoxy-04Spires et al., 2005; Garcia-Alloza et al., 2007
Mitochondrial membrane potentialLiverTetramethylrhodamine methyl ester Rhodamine 123Theruvath et al., 2008; Zhong et al., 2008
Oxygen consumptionLiverRu(phen3)2+Paxian et al., 2004
Sarcomere contraction in humansSkeletal muscleEndogenous fluorescenceLlewellyn et al., 2008
Open in a separate windowOpen in a separate windowFigure 3.Imaging tissues and individual cells in live animals. (A) The vasculature of an immunocompromised mouse was highlighted by the systemic injection of 2 MD dextran (red) before (left) and after (right) the implant of breast cancer cells in the back (green). Note the change in shape of the blood vessels and their increased permeability (arrow). Images were acquired by two-photon microscopy (excitation wavelength: 930 nm). (B) The microvasculature in the liver of a mouse expressing the membrane marker mTomato (red) was highlighted by the injection of cascade blue dextran (blue) and imaged by confocal microscopy (excitation wavelengths: 405 nm and 561 nm). Note the red blood cells that do not uptake the dye and appear as dark objects in the blood stream (arrow). (C) Metastatic and nonmetastatic human adenocarcinoma cells were injected in the tongue of an immunocompromised mouse and imaged for four consecutive days by using two-photon microscopy (excitation wavelength: 930 nm). The metastatic cells, which express the fluorescent protein mCherry (red), migrate away from the edge of the tumor (arrows), whereas the nonmetastatic cells, which express the fluorescent protein Venus (green), do not. (D) A granulocyte moving inside a blood vessel in the mammary gland of a mouse expressing GFP-tagged myosin IIb (green) and labeled with the mitochondrial vital dye MitoTracker (red) was imaged in time lapse by using confocal microscopy (excitation wavelengths: 488 nm and 561 nm). Figure corresponds to Video 1. Time is expressed as minutes:seconds. Bars: (A) 100 µm; (B) 10 µm; (C) 30 µm; (D) 10 µm.IVM has also been used successfully to study the dynamics and the morphological changes of individual cells within a tissue (
EventOrganProbeReference
Neuronal morphology of hippocampal neuronsBrainThy1-GFP mice, dextranBarretto et al., 2011
Neuronal circuitryBrainBrainbow miceLivet et al., 2007
Dendritic spine development in the cortexBrainYFP H-line micePan and Gan, 2008
Calcium imaging in the brainBrainGCAMPZariwala et al., 2012
Natural killer cell and cytotoxic T cell interactions with tumorsXenograftmCFP , mYFPDeguine et al., 2010
Neutrophil recruitment in beating heartHeartDextran, CX3CR1-GFP miceLi et al., 2012
Immune cells in the central nervous systemBrainDextran, CX3CR1-GFP, LysM-GFP and CD11c-YFP miceNayak et al., 2012
Dendritic cells migrationSkinYFP, VE-caherin RFP mice, dextranNitschké et al., 2012
CD8+ T cells interaction with dendritic cells during viral infectionLymph nodesEGFP, Dextran, SHGHickman et al., 2008
B cells and dendritic cells interactions outside lymph nodesLymph nodesEGFPQi et al., 2006
Change in gene expression during metastasisXenograftPinner et al., 2009
Invasion and metastasis in head and neck cancerXenograftYFP, RFP-lifeact, dextranAmornphimoltham et al., 2013
Fibrosarcoma cell migration along collagen fibersDorsal skin chamberSHG, EGFP, DsRed, DextranAlexander et al., 2008
Long term imaging mammary tumors and photo-switchable probesMammary windowDendra-2Kedrin et al., 2008; Gligorijevic et al., 2009
Long term imaging liver metastasis through abdominal windowLiverSHG, Dendra2, EGFPRitsma et al., 2012b
Macrophages during intravasation in mammary tumorsXenograftEGFP, SHG, dextransWang et al., 2007; Wyckoff et al., 2007
Melanoma collective migrationDorsal skin ChamberSHG, THG, EGFP, DextranWeigelin et al., 2012
Hematopoietic stem cells and blood vesselSkullcupDextranLo Celso et al., 2009
Epithelial stem cells during hair regenerationSkinH2B-GFP miceRompolas et al., 2012
Open in a separate windowIn neurobiology, for example, the development of approaches to perform long-term in vivo imaging has permitted the correlation of changes in neuronal morphology and neuronal circuitry to pathological conditions such as stroke (Zhang and Murphy, 2007), tumors (Barretto et al., 2011), neurodegenerative diseases (Merlini et al., 2012), and infections (McGavern and Kang, 2011). This has been accomplished by the establishment of surgical procedures to expose the brain cortex, and the implantation of chronic ports of observations such as cranial windows and imaging guide tubes for micro-optical probes (Svoboda and Yasuda, 2006; Xu et al., 2007; Barretto et al., 2011). In addition, this field has thrived thanks to the development of several transgenic mouse models harboring specific neuronal populations expressing either one or multiple fluorescent molecules (Svoboda and Yasuda, 2006; Livet et al., 2007).In tumor biology, the ability to visualize the motility of cancer cells within a tumor in vivo has provided tremendous information on the mechanisms regulating invasion and metastasis (Fig. 3 C; Beerling et al., 2011). Tumor cells metastasize to distal sites by using a combination of processes, which include tumor outgrowth, vascular intravasation, lymphatic invasion, or migration along components of the extracellular matrix and nerve fibers. Although classical histological analysis and indirect immunofluorescence have been routinely used to study these processes, the ability to perform long-term IVM through the optimization of optical windows (Alexander et al., 2008; Kedrin et al., 2008; Gligorijevic et al., 2009; Ritsma et al., 2012b) has provided unique insights. For example, a longitudinal study performed by using a combination of two-photon microscopy, SHG, and THG has highlighted the fact that various tissue components associated with melanomas may play either a migration-enhancing or migration-impeding role during collective cell invasion (Weigelin et al., 2012). In mammary tumors, the intravasation of metastatic cells has been shown to require macrophages (Wang et al., 2007; Wyckoff et al., 2007). In head and neck cancer, cells have been shown to migrate from specific sites at the edge of the tumor, and to colonize the cervical lymph nodes by migrating though the lymphatic vessels (Fig. 3 C; Amornphimoltham et al., 2013). In highly invasive melanomas, the migratory ability of cells has been correlated with their differentiation state, as determined by the expression of a reporter for melanin expression (Pinner et al., 2009).Imaging the cells of the immune system in a live animal has revealed novel qualitative and quantitative aspects of the dynamics of cellular immunity (Fig. 2 C and Video 1; Germain et al., 2005; Cahalan and Parker, 2008; Nitschke et al., 2008). Indeed, the very complex nature of the immune response, the involvement of a multitude of tissue components, and its tight spatial and temporal coordination clearly indicate that IVM is the most suited approach to study cellular immunity. This is highlighted in studies either in lymphoid tissues, where the exquisite coordination between cell–cell interactions and cell signaling has been studied during the interactions of B lymphocytes and T cell lymphoid tissues (Qi et al., 2006), T cell activation (Hickman et al., 2008; Friedman et al., 2010), and migration of dendritic cells (Nitschké et al., 2012), or outside lymphoid tissues, such as, for example, brain during pathogen infections (Nayak et al., 2012), heart during inflammation (Li et al., 2012), and solid tumors (Deguine et al., 2010).

Imaging subcellular structures in vivo and its application to cell biology

The examples described so far convey that IVM has contributed to unraveling how the unique properties of the tissue environment in vivo significantly regulate the dynamics of individual cells and ultimately tissue physiology. Is IVM suitable to determine (1) how subcellular events occur in vivo, (2) whether they differ in in vitro settings, and (3), finally, the nature of their contribution to tissue physiology?IVM has been extensively used to image subcellular structures in smaller organisms (i.e., zebrafish, Caenorhabditis elegans) that are transparent and can be easily immobilized (Rohde and Yanik, 2011; Tserevelakis et al., 2011; Hove and Craig, 2012). In addition, the ability to easily perform genetic manipulations has made these systems extremely attractive to study several aspects of developmental and cell biology. However, their differences in term of organ physiology with respect to rodents do not make them suitable models for human diseases. For a long time, subcellular imaging in live rodents has been hampered by the motion artifacts derived from the heartbeat and respiration. Indeed, small shifts along the three axes make it practically impossible to visualize structures whose sizes are in the micrometer or submicrometer range, whereas it marginally affects larger structures. This issue has been only recently addressed by using a combination of strategies, which include: (1) the development of specific surgical procedures that allow the exposure and proper positioning of the organ of interest (Masedunskas et al., 2013), (2) the improvement of specific organ holders (Cao et al., 2012; Masedunskas et al., 2012a), and (3) the synchronization of the imaging acquisition with the heartbeat and respiration (Presson et al., 2011; Li et al., 2012). Very importantly, these approaches have been successfully implemented without compromising the integrity and the physiology of the tissue, thus opening the door to study cell biology in a live animal.For example, large subcellular structures such as the nuclei have been easily imaged, making it possible to study processes such as cell division and apoptosis (Fig. 4 A; Goetz et al., 2011; Orth et al., 2011; Rompolas et al., 2012). Interestingly, these studies have highlighted the fact that the in vivo microenvironment substantially affects nuclear dynamics. Indeed, mitosis and the structure of the mitotic spindle were followed over time in a xenograft model of human cancer expressing the histone marker mCherry-H2B and GFP-tubulin (Orth et al., 2011). Specifically, the effects of the anticancer drug Paclitaxel were studied, revealing that the tumor cells in vivo have a higher mitotic index and lower pro-apoptotic propensity than in vitro (Orth et al., 2011). FRET has been used in subcutaneous tumors to image cytotoxic T lymphocyte–induced apoptosis and highlighted that the kinetics of this process are much slower than those reported for nontumor cells in vivo that are exposed to a different microenvironment (Breart et al., 2008). Cell division has also been followed in the hair-follicle stem cells of transgenic mice expressing GFP-H2B. This study determined that epithelial–mesenchymal interactions are essential for stem cell activation and regeneration, and that nuclear divisions occur in a specific area of the hair follicles and are oriented toward the axis of growth (Rompolas et al., 2012). These processes show an extremely high level of temporal and spatial organization that can only be appreciated in vivo and by using time-lapse imaging.Open in a separate windowFigure 4.Imaging subcellular events in live animals. (A) Human squamous carcinoma cells were engineered to stably express the Fucci cell cycle reporter into the nucleus and injected in the back of an immunocompromised mouse. After 1 wk, the tumor was imaged by two-photon microscopy and SHG (excitation wavelength: 930 nm). (top) Maximal projection of a z stack (xy view). Cells in G2/M are in green, cells in G1 are in red, and collagen fibers are in cyan. (bottom) Lateral view (xz) of a z stack. (B) Clusters of GLUT4-containing vesicles (green) in the soleus muscle of a transgenic mouse expressing GFP-GLUT4 and injected with 70 kD Texas red–dextran to visualize the vasculature and imaged by two-photon microscopy (excitation wavelength: 930 nm). (C) Confocal microscopy (excitation wavelength: 488 nm) of hepatocytes in the liver of a transgenic mouse expressing the autophagy marker GFP-LC3. The inset shows small GFP-LC3 autophagic vesicles. (D–G) Dynamics of intracellular compartments imaged by time-lapse two-photon (E) or confocal microscopy (D, F, and G). (D) Endocytosis of systemically injected 10 kD Texas red–dextran into the kidney of a transgenic mouse expressing the membrane marker m-GFP. The dextran (red) is transported from the microvasculature into the proximal tubuli, and then internalized in small endocytic vesicles (arrows; Video 2). (E) Endocytosis of a systemically injected 10 kD of Alexa Fluor 488 dextran into the salivary glands of a live rat. The dextran (green) diffuses from the vasculature into the stroma, and it is internalized by stromal cells (insets). Collagen fibers (red) are highlighted by SHG. (F) Regulated exocytosis of large secretory granules in the salivary glands of a live transgenic mouse expressing cytoplasmic GFP. The GFP is excluded from the secretory granules and accumulates on their limiting membranes (arrows) after fusion with the plasma membrane (broken lines). The gradual collapse of an individual granule is highlighted in the insets. (G) Dynamics of mitochondria labeled with the membrane potential dye TMRM in the salivary glands of a live mouse. Time is expressed as minutes:seconds. Bars: (A) 40 µm; (B) 15 µm; (C, D, E, and G) 10 µm; (F) 5 µm.Imaging membrane trafficking has been more challenging because of its dynamic nature and the size of the structures to image. The first successful attempt to visualize membrane traffic events was achieved in the kidney of live rats by using two-photon microscopy where the endocytosis of fluid-phase markers, such as dextrans, or the receptor-mediated uptake of folate, albumin, and the aminoglycoside gentamicin were followed in the proximal tubuli (Fig. 4 D and Video 2; Dunn et al., 2002; Sandoval et al., 2004; Russo et al., 2007). These pioneering studies showed for the first time that apical uptake is involved in the filtration of large molecules in the kidney, whereas previously it was believed to be exclusively due to a barrier in the glomerular capillary wall. However, in the kidney the residual motion artifacts limited the imaging to short periods of time. Recently, the salivary glands have proven to be a suitable organ to study the dynamics of membrane trafficking by using either two-photon or confocal microscopy. Systemically injected dextrans, BSA, and transferrin were observed to rapidly internalize in the stromal cells surrounding the salivary gland epithelium in a process dependent on the actin cytoskeleton (Masedunskas and Weigert, 2008; Masedunskas et al., 2012b). Moreover, the trafficking of these molecules through the endo-lysosomal system was documented, providing interesting insights on early endosomal fusion (Fig. 4 E; Masedunskas and Weigert, 2008; Masedunskas et al., 2012b). Notably, significant differences were observed in the kinetics of internalization of transferrin and dextran. In vivo, dextran was rapidly internalized by stromal cells, whereas transferrin appeared in endosomal structures after 10–15 min. However, in freshly explanted stromal cells adherent on glass, transferrin was internalized within 1 min, whereas dextran appeared in endosomal structures after 10–15 min. Although the reasons for this difference were not addressed, it is clear that the environment in vivo has profound effects on the regulation of intracellular processes (Masedunskas et al., 2012b). Similar differences have been reported for the caveolae that in vivo are more dynamic than in cell cultures (Thomsen et al., 2002; Oh et al., 2007). Endocytosis has also been investigated in the epithelium of the salivary glands (Sramkova et al., 2009). Specifically, plasmid DNA was shown to be internalized by a clathrin-independent pathway from the apical plasma membrane of acinar and ductal cells, and to subsequently escape from the endo-lysosomal system, thus providing useful information on the mechanisms of nonviral gene delivery in vivo (Sramkova et al., 2012). Receptor-mediated endocytosis has also been studied in cancer models. Indeed, the uptake of a fluorescent EGF conjugated to carbon nanotubes has been followed in xenografts of head and neck cancer cells revealing that the internalization occurs primarily in cells that express high levels of EGFR (Bhirde et al., 2009). The role of endosomal recycling has also been investigated during tumor progression. Indeed, the small GTPase Rab25 was found to regulate the ability of head neck cancer cells to migrate to lymph nodes by controlling the dynamic assembly of plasma membrane actin reach protrusion in vivo (Amornphimoltham et al., 2013). Interestingly, this activity of Rab25 was reconstituted in cells migrating through a 3D collagen matrix but not in cells grown adherent to a solid substrate.IVM has been a powerful tool in investigating the molecular machinery controlling regulated exocytosis in various organs. In salivary glands, the use of selected transgenic mice expressing either soluble GFP or a membrane-targeted peptide has permitted the characterization of the dynamics of exocytosis of the secretory granules after fusion with the plasma membrane (Fig. 4 E; Masedunskas et al., 2011a, 2012d). These studies revealed that the regulation and the modality of exocytosis differ between in vivo and in vitro systems. Indeed, in vivo, regulated exocytosis is controlled by stimulation of the β-adrenergic receptor, and secretory granules undergo a gradual collapse after fusion with the apical plasma membrane, whereas, in vitro, regulated exocytosis is also controlled by the muscarinic receptor and the secretory granules fuse to each other, forming strings of interconnected vesicles at the plasma membrane (compound exocytosis; Masedunskas et al., 2011a, 2012d). Moreover, the transient expression of reporter molecules for F-actin has revealed the requirement for the assembly of an actomyosin complex to facilitate the completion of the exocytic process (Masedunskas et al., 2011a, 2012d). This result underscores the fact that the dynamics of the assembly of the actin cytoskeleton can be studied both qualitatively and quantitatively in live animals at the level of individual secretory granules. In addition, this approach has highlighted some of the mechanisms that contribute to regulate the apical plasma membrane homeostasis in vivo that cannot be recapitulated in an in vitro model systems (Masedunskas et al., 2011b, 2012c; Porat-Shliom et al., 2013). Indeed, the hydrostatic pressure that is built inside the ductal system by the secretion of fluids that accompanies exocytosis plays a significant role in controlling the dynamics of secretory granules at the apical plasma membrane. This aspect has never been appreciated in organ explants where the integrity of the ductal system is compromised. Finally, a very promising model has been developed in the skeletal muscle, where the transient transfection of a GFP-tagged version of the glucose transporter type 4 (GLUT4) has made possible to characterize the kinetics of the GLUT4-containing vesicles in resting conditions and their insulin-dependent translocation to the plasma membrane (Fig. 4 B; Lauritzen et al., 2008, 2010). This represents a very powerful experimental model that bridges together physiology and cell biology and has the potential to provide fundamental information on metabolic diseases.These examples underscore the merits of subcellular IVM to investigate specific areas of cell biology such as membrane trafficking, the cell cycle, apoptosis, and cytoskeletal organization. However, IVM is rapidly extending to other areas, such as cell signaling (Stockholm et al., 2005; Rudolf et al., 2006; Ritsma et al., 2012a), metabolism (Fig. 4 C; Débarre et al., 2006; Cao et al., 2012), mitochondrial dynamics (Fig. 4 F; Sun et al., 2005; Hall et al., 2013), or gene and protein expression (Pinner et al., 2009) that have just begun to be explored.

Future perspectives

IVM has become a powerful tool to study biological processes in live animals that is destined to have an enormous impact on cell biology. The examples described here give a clear picture of the broad applicability of this approach. In essence, we foresee that IVM is going to be the obligatory choice to study highly dynamic subcellular processes that cannot be reconstituted in vitro or ex vivo, or when a link between cellular events and tissue physiopathology is being pursued. In addition, IVM will provide the opportunity to complement and confirm data generated from in vitro studies. Importantly, the fact that in several instances confocal microscopy can be effectively used for subcellular IVM makes this approach immediately accessible to several investigators.In terms of future directions, we envision that other light microscopy techniques will soon become standard tools for in vivo studies, as shown by the recent application of FRET to study signaling (Stockholm et al., 2005; Rudolf et al., 2006; Breart et al., 2008; Ritsma et al., 2012a), and FRAP, which has been used in the live brain to measure the diffusion of α synuclein, thus opening the door to studying the biophysical properties of proteins in vivo (Unni et al., 2010). Moreover, super-resolution microscopy may be applied for imaging live animals, although this task may pose some challenges. Indeed, these techniques require: (1) the complete stability of the specimen, (2) extended periods of time for light collection, (3) substantial modifications to the existing microscopes, and (4) the generation of transgenic mice expressing photoactivatable probes.To reach its full potential, IVM has to further develop two main aspects: animal models and instrumentations. Indeed, a significant effort has to be invested in developing novel transgenic mouse models, which express fluorescently labeled reporter molecules. One example is the recently developed mouse that expresses fluorescently tagged lifeact. This model will provide the unique opportunity to study F-actin dynamics in vivo in the context of processes such as cell migration and membrane trafficking (Riedl et al., 2010). Moreover, the possibility of crossing these reporter mice with knockout animals will provide the means to further study cellular processes at a molecular level. Alternatively, reporter molecules or other transgenes that may perturb a specific cellular pathway can be transiently transfected into live animals in several ways. Indeed, the remarkable advancements in gene therapy have contributed to the development of several nonviral- and viral-mediated strategies for gene delivery to selected target organs. In this respect, the salivary glands and the skeletal muscle are two formidable model systems because either transgenes or siRNAs can be successfully delivered without any adverse reaction and expressed in a few hours. In terms of the current technical limitations of IVM, the main areas of improvement are the temporal resolution, the ability to access the organ of interest with minimal invasion, and the ability to perform long-term imaging. As for the temporal resolution, the issue has begun to be addressed by using two different approaches: (1) the use of spinning disk microscopy, as shown by its recent application to image platelet dynamics in live mice (Jenne et al., 2011); and (2) the development of confocal and two-photon microscopes equipped with resonant scanners that permit increasing the scanning speed to 30 frames per second (Kirkpatrick et al., 2012). As for accessing the organs, recently several microlenses (350 µm in diameter) have been inserted or permanently implanted into live animals, minimizing the exposure of the organs and the risk of affecting their physiology (Llewellyn et al., 2008). Finally, although some approaches for the long-term imaging of the brain, the mammary glands, and the liver have been developed, additional effort has to be devoted to establish chronic ports of observations in other organs.In conclusion, these are truly exciting times, and a new era full of novel discoveries is just around the corner. The ability to see processes inside the cells of a live animal is no longer a dream.

Online supplemental material

Video 1 shows time-lapse confocal microscopy of a granulocyte moving inside a blood vessel in the mammary gland of a mouse expressing GFP-tagged myosin IIb (green) and labeled with MitoTracker (red). Video 2 shows time-lapse confocal microscopy of the endocytosis of systemically injected 10 kD Texas red–dextran (red) into the kidney-proximal tubuli of a transgenic mouse expressing the membrane marker m-GFP (green). Online supplemental material is available at http://www.jcb.org/cgi/content/full/jcb.201212130/DC1.  相似文献   

10.
Quality control: Organellophagy: Eliminating cellular building blocks via selective autophagy     
Koji Okamoto 《The Journal of cell biology》2014,205(4):435-445
Maintenance of organellar quality and quantity is critical for cellular homeostasis and adaptation to variable environments. Emerging evidence demonstrates that this kind of control is achieved by selective elimination of organelles via autophagy, termed organellophagy. Organellophagy consists of three key steps: induction, cargo tagging, and sequestration, which involve signaling pathways, organellar landmark molecules, and core autophagy-related proteins, respectively. In addition, posttranslational modifications such as phosphorylation and ubiquitination play important roles in recruiting and tailoring the autophagy machinery to each organelle. The basic principles underlying organellophagy are conserved from yeast to mammals, highlighting its biological relevance in eukaryotic cells.

Introduction

Organelles are fundamental subunits of eukaryotic cells that possess structurally and functionally distinct characteristics that allow them to perform unique activities crucial for viability. It is thus a matter of the utmost importance for cells to maintain organellar quality and integrity. In addition, cells modulate the quantity of organelles in order to balance organellar activities and cellular demands, which can act as an adaptive mechanism to diverse environmental changes. Both dysfunctional and surplus organelles are cleared from cells through autophagy, a widely conserved self-eating process by which cytoplasmic constituents are sequestered as cargoes by intracellular membranes that fuse with lysosomes for hydrolytic breakdown (Mizushima and Komatsu, 2011; Mizushima et al., 2011; Weidberg et al., 2011).Although autophagy has primarily been recognized as a nonselective degradation pathway, recent studies reveal that it also plays a vital role in digesting specific cargoes such as proteins and organelles (Mizushima, 2011; Suzuki, 2013). The latter process, termed selective autophagy, includes the following three critical stages: first, signaling from degradation cues induces downstream events specific for a particular target; second, regulation of landmark molecules that tag the target as disposable cargo; third, assembly of core autophagy-related (Atg) proteins to sequester the cargo. In many cases, malfunction in or decreased cellular metabolism related to a protein or organelle leads to expression and activation of a landmark molecule. Core Atg proteins then localize to the cargo via direct or indirect interactions with the landmark molecule and ultimately mediate selective autophagy.In this short review, we summarize recent findings on organellophagy, autophagy-related pathways selective for organelles such as the peroxisome, mitochondrion, lipid droplet (structure surrounded by a phospholipid monolayer), lysosome, nucleus, ER, and even nonmembraneous structures like the ribosome. Despite the diversity of their degradation cues and landmark molecules, organellophagy seems to be regulated by common basic principles involving protein phosphorylation and ubiquitination. In particular, these two major posttranslational modifications promote targeting of core Atg proteins to the organellar surface. Defects in several organellophagy pathways are associated with various disorders including renal injury, neurodegeneration, obesity, and atherosclerosis (Mizushima and Komatsu, 2011), underscoring their physiological significance in health and disease.

Modes of autophagy in organellophagy

Three morphologically distinct modes of autophagic processes have so far been defined: macroautophagy, microautophagy, and chaperone-mediated autophagy (CMA; Fig. 1; Mizushima and Komatsu, 2011; Li et al., 2012; Cuervo and Wong, 2014; Feng et al., 2014). Macro- and microautophagy are conserved from yeast to humans, whereas CMA has been found only in mammals. Upon macroautophagy induction, newly formed double membrane–bound structures enclose proteins and organelles, eventually generating mature vesicles, called autophagosomes. Core Atg proteins play essential roles in autophagosome formation. The engulfed cargoes are then mixed with the lysosomal hydrolases via autophagosome–lysosome fusion and digested into small molecules for recycling. During microautophagy, the lysosomal membrane invaginates to sequester proteins and organelles. In some cases, accompanying membrane structures function in closure of the cargoes, which requires core Atg proteins. The lysosomal lipase then digests the internalized vesicles, leading to breakdown of the cargoes by hydrolases. By contrast, CMA recruits specific protein substrates associated with the molecular chaperone Hsc70 to lysosomes and translocates the substrates one by one into the lysosomal lumen through the receptor protein Lamp-2A in a manner independent on core Atg proteins. Unlike macro- and microautophagy, CMA has been suggested to degrade only proteins but not whole organelles.Open in a separate windowFigure 1.Three distinct modes of autophagy. In macroautophagy, newly generated cup-shaped structures, called isolation membranes, expand to surround cytoplasmic components. The two edges of isolation membranes then fuse to form double membrane–bound autophagosomes. Subsequently, autophagosomes fuse to lysosomes, and the engulfed cargoes are digested by hydrolytic enzymes. In microautophagy, invagination of the lysosomal membrane occurs to sequester proteins and organelles in the cytosol. The resulting vesicular structures are then pinched off and released into the lysosomal lumen for digestion. In chaperone-mediated autophagy (CMA), the Hsc70/co-chaperone complex delivers specific substrate proteins to lysosomes. The substrate polypeptides are then translocated one by one through the lysosomal membrane protein Lamp-2A and digested in the lysosomal lumen. Macro- and microautophagy are conserved from yeast to humans, whereas CMA has been found only in mammals. Unlike macro- and microautophagy, CMA has been suggested to degrade only proteins but not whole organelles.Morphological classification of organelle-specific autophagy in yeast and mammals is summarized in Manjithaya et al., 2010b; Oku and Sakai, 2010), mitochondrion (mitophagy; Ashrafi and Schwarz, 2013; Feng et al., 2013), lipid droplet (lipophagy; Liu and Czaja, 2013), and nucleus (nucleophagy; Mijaljica and Devenish, 2013). These degradation pathways appear to be conserved from yeast to humans. Macroautophagy-related turnover processes specific for lysosome (lysophagy; Hung et al., 2013; Maejima et al., 2013) and ER (reticulophagy/ER-phagy; Bernales et al., 2006) have been found in mammals and yeast, respectively. Although whether ribosome degradation in yeast occurs via macro- or microautophagy remains to be clarified, it seems to be a selective event (ribophagy) because ribosomal subunits are degraded significantly faster than other cytosolic proteins in an autophagy-dependent fashion (Kraft et al., 2008). To date, there has been no evidence suggesting selective degradation of the Golgi apparatus (golgiphagy).

Table 1.

Classification of organellophagy
Cargo organelleMacroautophagyMicroautophagy
YeastHumanYeastHuman
PeroxisomeND
MitochondrionND
Lipid dropletNDND
NucleusNDND
LysosomeNDNDND
Endoplasmic reticulumNDNDND
RibosomeaNDNDND
Open in a separate windowND, not determined.aRibophagy seems to depend on macroautophagy rather than microautophagy, although this has not yet been confirmed morphologically.

Common features of organellophagy

Studies on pexophagy and mitophagy have extensively explored molecular mechanisms underlying cargo recognition, implicating two common types, receptor- and ubiquitin-mediated processes (Fig. 2). Both types involve protein phosphorylation that activates or inactivates downstream events.Open in a separate windowFigure 2.Two common mechanisms of organellophagy. Molecular mechanisms underlying cargo recognition in pexophagy and mitophagy have extensively been explored, including two common types, receptor- and ubiquitin-mediated processes. Both types involve protein phosphorylation that activates or inactivates their downstream events. In the receptor-mediated process, membrane-anchored or peripherally associated receptors on the organellar surface interact with Atg8/LC3, ubiquitin-like proteins conjugated to the phospholipid phosphatidylethanolamine and localized to autophagosomes, and Atg11/Atg17, scaffold proteins required for core Atg protein assembly. Protein kinases phosphorylate receptors and regulate receptor interactions with Atg8/LC3 and Atg11/Atg17. In the ubiquitin-mediated process, E3 ubiquitin ligases target to the organelle and ubiquitinate proteins on the organellar surface. The ubiquitin chains then interact with LC3-binding adaptors such as p62/NBR1, or unknown factors (X) that may promote core Atg protein assembly. Protein kinases phosphorylate the ubiquitin ligases and promote targeting and activation of the E3 enzymes.In the receptor-mediated process, specific proteins membrane-anchored or tightly associated on the organellar surface interact directly, or indirectly via adaptor proteins, with Atg8 (LC3, GABARAP, and GATE-16 in mammalian cells), a highly conserved ubiquitin-like protein essential for all autophagy-related pathways (Shpilka et al., 2011; Rogov et al., 2014; Wild et al., 2014). Notably, receptor proteins contain tetrapeptide consensus sequences called Atg8 family–interacting motif (AIM) and LC3-interacting region (LIR) that consist of W/YxxI/L/V and W/F/YxxL/I/V, respectively (Noda et al., 2010; Birgisdottir et al., 2013). AIM/LIR directly associates with Atg8/LC3 through the side chains of their conserved residues bound deeply into the hydrophobic pocket of Atg8/LC3. Mutations in AIM and LIR impair degradation of cargo organelles, suggesting the significance of these interactions. Because Atg8 is covalently linked to the phospholipid phosphatidylethanolamine and localized predominantly to autophagosomes, the receptor–Atg8/LC3 interactions could assist generation and expansion of cup-shaped structures called isolation membranes surrounding cargo organelles. In yeast, pexophagy and mitophagy receptors also interact with Atg11 or Atg17, scaffold proteins that serve as platforms for core Atg protein assembly (Farré et al., 2008; Kanki et al., 2009; Okamoto et al., 2009; Motley et al., 2012). Importantly, protein kinases and phosphatases modify receptors and appear to play regulatory roles in stabilizing or destabilizing the interactions of receptor proteins with Atg8/LC3 and Atg11/Atg17 (Farré et al., 2008, 2013; Novak et al., 2010; Aoki et al., 2011; Kondo-Okamoto et al., 2012; Liu et al., 2012; Kanki et al., 2013; Zhu et al., 2013).In the ubiquitin-mediated process, peripheral and/or membrane-anchored proteins on the surface of cargo organelles are ubiquitinated by specific E3 ligases (Shaid et al., 2013). These ubiquitin chains act as “degradation tags” recognized by soluble adaptor proteins such as p62 and NBR1 that also interact with LC3 (Johansen and Lamark, 2011). Targeting of other core Atg proteins to these cargo organelles seems to be independent of p62 and LC3 (Itakura et al., 2012), which may be mediated directly by ubiquitin, or indirectly via unknown ubiquitin-binding proteins. In some cases, mitophagy-specific E3 ligases are regulated by phosphorylation. For example, the protein kinase PINK1 phosphorylates the ubiquitin E3 ligase Parkin to promote mitophagy (Kondapalli et al., 2012; Shiba-Fukushima et al., 2012; Iguchi et al., 2013). This type of mitophagy has so far been found in mammals, but not in yeast. Finally, it should be noted that the receptor- and ubiquitin-mediated processes are not mutually exclusive, as the LC3 receptor Pex14 is also involved in the ubiquitin/NBR1-mediated pexophagy in mammalian cells (Deosaran et al., 2013).

Pexophagy

In response to changes in the intra- and extracellular environments, peroxisome number dynamically increases or decreases in order to maintain the appropriate levels of the metabolic reactions including fatty acid oxidation and H2O2 detoxification (Smith and Aitchison, 2013). For example, the methylotrophic yeasts Pichia pastoris and Hansenula polymorpha can proliferate large peroxisome clusters when they grow in media containing methanol as the sole carbon source (van der Klei et al., 2006). Pexophagy is then drastically triggered upon a shift from methanol to glucose or ethanol media in which the peroxisomal metabolism is not critical for cell growth and viability (Manjithaya et al., 2010b; Oku and Sakai, 2010). Thus, molecular mechanisms underlying the selectivity of pexophagy have mostly been uncovered in these methylotrophic yeasts.Macro- and microautophagy mediate pexophagy (macro- and micropexophagy, respectively) in P. pastoris that requires the soluble receptor protein Atg30 that interacts with Pex3 and Pex14, two peroxisomal membrane proteins, and recruits Atg8, Atg11, and Atg17 to the surface of peroxisomes (Farré et al., 2008, 2013). Similarly, Atg36 acts as a soluble receptor protein in the budding yeast Saccharomyces cerevisiae, localizes to the peroxisomal surface via Pex3, and binds Atg8 and Atg11 to promote macropexophagy (Motley et al., 2012; Farré et al., 2013). Interestingly, both Atg30 and Atg36 contain AIMs flanked with putative phosphoserine residues (Farré et al., 2013). These amino acids modified by unknown kinase(s) may stabilize the Atg30–Atg8 and Atg36–Atg8 interactions. Additional phosphorylation sites are also required for binding of Atg30 and Atg36 to Atg11 (Farré et al., 2008, 2013). In S. cerevisiae, the MAPK cascade Mid2–Pkc1–Bck1–Mkk1/Mkk2–Slt2 is necessary for peroxisome degradation, but not for pexophagosome formation (Manjithaya et al., 2010a; Mao et al., 2011). Hence, Slt2 is unlikely to regulate the interactions of Atg36 with Atg8 and Atg11. Nonetheless, pexophagy in S. cerevisiae is enhanced in a set of mutants containing dysfunctional peroxisomes through yet-uncharacterized modifications of Atg36 (Nuttall et al., 2014). Despite the common role in recruiting the pexophagy receptors to the peroxisomal surface in P. pastoris and S. cerevisiae, Pex3 in H. polymorpha is degraded via the ubiquitin–proteasome pathway in order to initiate macropexophagy by unknown mechanisms (Bellu et al., 2002; Williams and van der Klei, 2013), suggesting the diversity of peroxisome turnover mechanisms among yeast species. It should also be noted that dynamin-related GTPases, Dnm1 and Vps1, target to peroxisomes, and promote peroxisomal fission, which is a critical step before pexophagy in H. polymorpha and S. cerevisiae (Manivannan et al., 2013; Mao et al., 2014).A study using Chinese hamster ovary cells demonstrates that pexophagy can be induced upon a shift from starvation to nutrient-rich media (Hara-Kuge and Fujiki, 2008). Under this condition, Pex14 interacts with LC3-II, a phosphatidylethanolamine-conjugated form anchored on autophagosomes (Hara-Kuge and Fujiki, 2008). When monoubiquitinated peroxisomal membrane proteins are overexpressed in COS-7 cells, pexophagy occurs in a manner dependent on p62 (Kim et al., 2008). More recently, down-regulation of either p62 or NBR1 has been shown to suppress degradation of peroxisomes in HeLa cells (Deosaran et al., 2013). Overexpression of NBR1, but not p62, can facilitate pexophagy through its LIR, coiled-coil domain (for homo-oligomerization), JUBA domain (for membrane association), and UBA domain (for ubiquitin binding; Deosaran et al., 2013). Notably, an NBR1 mutant defective in p62 interaction is not fully functional for pexophagy (Deosaran et al., 2013). Thus, p62 may not be a major adaptor, but still contributes to pexophagy in cooperation with NBR1. Nonetheless, it seems likely that ubiquitination of peroxisomal proteins promotes recruitment of LC3 to peroxisomes via p62 and NBR1, ultimately leading to pexophagy in mammalian cells. How the core factors of the autophagy machinery are targeted to peroxisomes remains to be clarified.

Mitophagy

Mitochondria are major organelles that are platforms for many important processes including energy conversion, calcium homeostasis, and programmed cell death (Nunnari and Suomalainen, 2012). These organelles concomitantly generate reactive oxygen species (ROS) as hazardous byproducts during respiration. Consequently, accumulation of ROS causes mitochondrial dysfunction. Elimination of damaged mitochondria is therefore critical for cell homeostasis (Okamoto and Kondo-Okamoto, 2012). The other problem related to their energy metabolism is that cells need to maintain the balance between ATP supply and demand. Upon a shift from high to low energy consumption state, surplus mitochondria become vital targets for clearance (Okamoto and Kondo-Okamoto, 2012). Numerous studies demonstrate that mitophagy contributes to mitochondrial quality and quantity control, and that its selectivity is established via common mechanisms (Youle and Narendra, 2011; Jin and Youle, 2012; Narendra et al., 2012; Ashrafi and Schwarz, 2013; Feng et al., 2013).In the yeast S. cerevisiae, the mitophagy receptor Atg32 is induced in response to oxidative stress and anchored on the surface of mitochondria with its N- and C-terminal regions exposed to the cytosol and mitochondrial intermembrane space (IMS), respectively (Fig. 3 A; Kanki et al., 2009; Okamoto et al., 2009). Atg32 contains an AIM near the N terminus that is embedded into the hydrophobic pocket of Atg8 (Okamoto et al., 2009; Kondo-Okamoto et al., 2012). The C-terminal coiled-coil domain of Atg11 physically associates with Atg32 via a consensus region following the AIM (Aoki et al., 2011). Notably, this Atg11-interacting region contains serine residues that appear to be modified directly by casein kinase-2 (CK2), a housekeeping protein kinase (Kanki et al., 2013). This posttranslational modification stabilizes Atg32–Atg11 interaction (Fig. 3 A; Aoki et al., 2011; Kondo-Okamoto et al., 2012). A recent study suggests that processing of the Atg32 C-terminal region by Yme1, a catalytic subunit of the mitochondrial inner membrane AAA protease facing the IMS, is important for Atg32–Atg11 interaction (Wang et al., 2013), yet the role of Yme1 in mitophagy is currently a matter of debate (Campbell and Thorsness, 1998; Welter et al., 2013). In addition to CK2, the MAPK cascades Wsc1–Pkc1–Bck1–Mkk1/2–Slt2 and Ssk1–Pbs2–Hog1 are important for mitophagy (Aoki et al., 2011; Mao et al., 2011). Phosphorylation of Atg32 depends on Hog1, but not Slt2, while Atg32 is not a substrate for Hog1 (Aoki et al., 2011). Atg1, a protein kinase essential for all autophagy-related processes, is also involved in Atg32 phosphorylation (Kondo-Okamoto et al., 2012), although the molecular function of this modification remains unclear.Open in a separate windowFigure 3.Models for mitophagy in yeast and mammalian cells. (A) Atg32-mediated mitophagy in S. cerevisiae. Under respiratory conditions, the mitophagy receptor Atg32 is induced in response to oxidative stress, targeted, and anchored to the mitochondrial surface. Atg32 recruits Atg8 and Atg11 to mitochondria via distinct domains. CK2 phosphorylates Atg32 to stabilize the interaction between Atg32 and Atg11. This tertiary complex and core Atg proteins cooperatively generate isolation membranes to sequester mitochondria. The protein kinases Slt2 and Hog1 are also critical for mitophagy in yeast, although their targets remain unknown. (B) FUNDC1-mediated mitophagy in mammals. Under normoxic conditions, the mitochondrial outer membrane protein FUNDC1 is phosphorylated by Src and CK2, thereby preventing LC3 binding. Upon hypoxia, the expression of Src is strongly suppressed, and the protein phosphatase PGAM5 dephosphorylates FUNDC1 and promotes LC3 binding. In addition, ULK1, a mammalian Atg1 kinase homologue, interacts with FUNDC1 and phosphorylates the mitophagy receptor. This posttranslational modification also stabilizes the interaction between FUNDC1 and LC3. (C) PINK1/Parkin-mediated mitophagy in mammals. When targeted to healthy mitochondria, PINK1 is partially translocated across the mitochondrial membranes, proteolytically processed, released back to the cytosol, and rapidly degraded. In cells containing damaged mitochondria, PINK1 is stalled in the outer membrane and associated with the TOM complex. Two molecules of PINK1 undergo self-activation via autophosphorylation. Active PINK1 then phosphorylates Parkin and stabilizes the E3 ligase on the surface of mitochondria. Mitochondria-associated Parkin promotes ubiquitination of multiple substrates, ultimately leading to LC3 and p62/NBR1 recruitment and core Atg protein assembly. Ubiquitin chains and these proteins are bridged by an unknown factor (X).Similar to Atg32-mediated mitophagy in yeast, three mitochondria-anchored receptors, NIX, BNIP3, and FUNDC1, promote autophagic degradation selective for mitochondria in mammalian cells (Schweers et al., 2007; Sandoval et al., 2008; Zhang et al., 2008; Liu et al., 2012). All three proteins contain the LIR consensus sequences that are important for their mitophagy activities (Novak et al., 2010; Liu et al., 2012; Zhu et al., 2013). NIX is highly induced during reticulocyte maturation and interacts with LC3 and GABARAP (Schweers et al., 2007; Novak et al., 2010). BNIP3 is strongly expressed in response to hypoxia and activated by reoxygenation (Zhang et al., 2008; Zhu et al., 2013). Notably, phosphorylation of serine residues near the BNIP3 LIR is crucial for LC3 and GATE-16 binding, and efficient mitophagy (Zhu et al., 2013). Kinases regulating the BNIP3 LIR are currently unknown. Although FUNDC1 is constitutively expressed under normoxic conditions, the tyrosine residue of the LIR is phosphorylated by the Src family kinase, which prevents LC3 binding and mitophagy (Fig. 3 B; Liu et al., 2012). Strikingly, hypoxia strongly suppresses Src expression, leading to dephosphorylation of the FUNDC1 LIR by unknown protein phosphatases, subsequent binding of LC3, and ultimate activation of mitophagy (Fig. 3 B; Liu et al., 2012). Furthermore, a serine residue near the LIR is phosphorylated by CK2 under normal conditions, and conversely dephosphorylated by the mitochondrial phosphatase PGAM5 upon hypoxic stress and mitochondrial membrane potential (ΔΨm) dissipation, leading to efficient LC3 binding and mitophagy activation (Fig. 3 B; Chen et al., 2014). Another recent study reveals that ULK1, a mammalian Atg1 kinase, targets to mitochondria via interaction with FUNDC1 and phosphorylates the mitophagy receptor to stabilize the FUNDC1–LC3 interaction (Fig. 3 B; Wu et al., 2014). It is not certain whether these mitophagy receptors could also recruit core Atg proteins to mitochondria.In addition to the receptor-driven pathways described above, mammalian cells use the ubiquitin-dependent processes to promote degradation of mitochondria. The best known is the mitophagy involving PINK1, a mitochondrial protein kinase, and Parkin, a cytosolic E3 ubiquitin ligase, two closely related causal factors for autosomal-recessive familial Parkinsonism (Kitada et al., 1998; Valente et al., 2004). When targeted to healthy mitochondria, PINK1 is partially translocated across the mitochondrial outer and inner membranes, cleaved by several enzymes including the matrix-localized mitochondrial-processing peptidase MPP and the inner membrane protease PARL, released back into the cytosol, and rapidly degraded by the proteasome via the N-end rule pathway (Fig. 3 C; Jin et al., 2010; Matsuda et al., 2010; Narendra et al., 2010b; Deas et al., 2011; Meissner et al., 2011; Shi et al., 2011; Greene et al., 2012; Yamano and Youle, 2013). In this situation, Parkin is dispersed throughout the cytosol as an inactive form and is not stably associated with mitochondria (Narendra et al., 2008, 2010b; Matsuda et al., 2010; Chaugule et al., 2011; Chew et al., 2011) (Fig. 3 C). As a result, mitophagy is mostly suppressed in normally respiring cells. Upon mitochondrial dysfunction such as ΔΨm dissipation, PINK1 is stalled in the outer membrane and anchored on the surface of mitochondria (Kawajiri et al., 2010; Matsuda et al., 2010; Narendra et al., 2010b; Rakovic et al., 2010). Subsequently, PINK1 forms a supermolecular complex together with the translocase of the outer membrane (TOM) components (Fig. 3 C; Lazarou et al., 2012; Okatsu et al., 2013). In this supermolecular complex, two molecules of PINK1 undergo intermolecular phosphorylation (Okatsu et al., 2013). PINK1 complex formation is correlated well with its autophosphorylation, which is prerequisite for recruitment of Parkin to damaged mitochondria (Okatsu et al., 2012, 2013). Through these processes, PINK1 becomes more active, efficiently phosphorylating a serine residue of the Parkin ubiquitin-like (Ubl) domain (Fig. 3 C; Kondapalli et al., 2012; Shiba-Fukushima et al., 2012; Iguchi et al., 2013). Phosphorylation of the Ubl domain probably induces a conformational change, at least to some extent, resulting in Parkin self-association and ubiquitin-thioester formation at the RING2 domain, which is essential for the E3 ligase activity (Chaugule et al., 2011; Iguchi et al., 2013; Lazarou et al., 2013; Spratt et al., 2013; Zheng and Hunter, 2013). Importantly, these PINK1-mediated events are consistent with the mechanisms of Parkin inactive–active state transition revealed by recent structural studies (Riley et al., 2013; Trempe et al., 2013; Wauer and Komander, 2013).Whether specific Parkin targets are required for mitophagy remains controversial (Geisler et al., 2010; Lee et al., 2010; Narendra et al., 2010a; Okatsu et al., 2010). A high-throughput analysis on the Parkin-dependent ubiquitylome demonstrates numerous targets on the surface of depolarized mitochondria including mitofusins (Mfns), large GTPases required for mitochondrial fusion (Sarraf et al., 2013). Parkin is responsible for degradation of mitofusins, preventing refusion of damaged mitochondria and assisting subsequent mitophagy (Gegg et al., 2010; Tanaka et al., 2010; Glauser et al., 2011; Rakovic et al., 2011). The role of Mfns in the PINK1/Parkin pathway is rather intricate, as it has been reported that Mfn2 serves as a Parkin receptor to promote mitochondrial degradation in mouse cardiomyocytes (Chen and Dorn, 2013). Notably, genome-wide siRNA screens uncover additional factors for PINK1/Parkin-mediated mitophagy, including TOMM7, a component of the TOM complex, as essential for stabilizing PINK1 on the outer membrane of depolarized mitochondria (Hasson et al., 2013). Rab GTPase-activating proteins have recently been shown to interact with Fis1, a tail-anchored protein, and LC3/GABARAP family members on the surface of mitochondria where they promote formation of autophagosomes by regulating Rab7 activity during PINK1/Parkin-mediated mitophagy (Yamano et al., 2014). Very recently, three studies demonstrate that PINK1 phosphorylates ubiquitin to activate Parkin in a manner similar to Parkin self-activation via the phosphorylated Ubl domain (Kane et al., 2014; Kazlauskaite et al., 2014; Koyano et al., 2014). Whether ubiquitin/LC3-binding adaptors such as p62 and NBR1 are necessary for the PINK1/Parkin pathway, and how core Atg proteins are recruited to damaged mitochondria remain inconclusive.In addition to Parkin, two ubiquitin E3 ligases, Gp78 and SMURF1, have been implicated in mammalian mitophagy. The Gp78-mediated process depends on Mfn1, but does not require Parkin (Fu et al., 2013). In contrast, SMURF1 is required for the PINK1/Parkin pathway (Orvedahl et al., 2011). Molecular mechanisms underlying Gp78 and SMURF1 functions have not yet been elucidated.

Lipophagy

Lipid droplets (LDs) consist of a core mainly containing triglycerides and sterol esters surrounded by a phospholipid monolayer and associated with various proteins. They are dynamic organelles that change their size and number in response to diverse conditions, and play key roles in lipid storage and metabolism (Walther and Farese, 2012). In addition to the cytosolic lipases, lysosomal hydrolases catabolize LDs that are transported via lipophagy (Liu and Czaja, 2013). In yeast, LDs are degraded through microautophagy (van Zutphen et al., 2014). By contrast, lipophagy occurs via macroautophagy in mouse hepatocytes and human enterocytes (Singh et al., 2009; Khaldoun et al., 2014). How the selectivity of lipophagy is established needs future studies.

Nucleophagy

Accumulating evidence suggests that portions of the nucleus, nucleus-derived components, or even a whole nucleus, are degraded by selective autophagy in a variety of eukaryotes (Mijaljica and Devenish, 2013). These processes, defined as nucleophagy, can be induced under starvation and other stress conditions such as DNA damage and cell cycle arrest (Mijaljica and Devenish, 2013). In the yeast S. cerevisiae, small teardrop-shaped parts of the nucleus are engulfed by the vacuole, a lytic organelle equivalent to the lysosome, at nucleus–vacuole (NV) junctions (Roberts et al., 2003). This event, termed piecemeal microautophagy of the nucleus (PMN), is induced soon after nutrient deprivation (Roberts et al., 2003). Formation of NV junctions requires Nvj1 in the nuclear envelope and Vac8 on the vacuolar membrane, two physically associated proteins that establish the vacuolar diffusion barrier, invaginate NV junctions, and generate PMN vesicles in a manner dependent on the vacuolar electrochemical gradient and lipid-modifying enzymes (Roberts et al., 2003; Dawaliby and Mayer, 2010). Atg11, Atg17, and other core Atg proteins are indispensable for PMN, as in the case of micropexophagy in the methylotrophic yeasts (Krick et al., 2008). After prolonged starvation, another type of nucleophagy also occurs through unknown mechanisms, which does not require Nvj1, Vac8, and Atg11 (Mijaljica et al., 2012).In mammals, LC3- and several core Atg-positive structures containing nuclear components accumulate in close proximity to the nucleus in cells from nuclear envelopathies (Park et al., 2009). In addition, micronuclei, small structures containing displaced chromosomes or chromosome fragments efficiently generated in cells expose to genotoxic stress, are degraded via autophagy (Rello-Varona et al., 2012). These autophagic micronuclei are p62 positive and exhibit signs of nuclear envelope degradation and DNA damage (Rello-Varona et al., 2012). It has also been suggested that LC3-positive micronuclei represent vesicles containing DNA that has not been repaired (Erenpreisa et al., 2012). Whether macro- and microautophagy could mediate nucleophagy in mammals and how the selectivity is established remain to be clarified.

Lysophagy

Lysosomes are acidic organelles highly enriched with hydrolytic enzymes that digest macromolecules delivered via the endocytic and autophagic pathways. Recent studies demonstrate that lysosomal rupture causes release of hydrolases into the cytosol, ultimately leading to destruction of intracellular structures and functions (Boya and Kroemer, 2008). It is therefore conceivable that cells must use surveillance and quality control systems for lysosomes. Indeed, emerging evidence reveals that damaged lysosomes are selectively sequestered by macroautophagy in mammalian cells (Hung et al., 2013; Maejima et al., 2013). Lysophagy seems to be a ubiquitin-mediated process involving LC3 and p62, which could contribute to recovery of lysosomal activities (Maejima et al., 2013). How ubiquitin and core Atg proteins selectively target to damaged lysosomes awaits further investigations.

Reticulophagy/ER-phagy

ER membranes are most abundant in many cell types, and their lumens serve as major factories for protein folding and modification. Although macroautophagy in yeast under starvation conditions can nonselectively sequester the ER together with other cytoplasmic constituents, ER components are more enriched than cytosolic proteins in autophagic bodies, suggesting a selective feature of this ER turnover (Hamasaki et al., 2005). Strikingly, when yeast cells are challenged with protein folding stress, ER membrane stacks are densely enclosed in autophagosome-like structures (Bernales et al., 2006). Recently, a Ypt/Rab GTPase module containing Atg11 has been reported to regulate reticulophagy/ER-phagy in yeast (Lipatova et al., 2013). These observations raise the possibility that ER turnover occurs via unknown selective mechanisms.

Ribophagy

In yeast, a hallmark of starvation-induced, nonselective macroautophagy is that autophagic bodies in the vacuolar lumen contain myriad ribosomes (Takeshige et al., 1992). However, under the same conditions, ribosomal subunits are degraded faster than other cytosolic proteins (Kraft et al., 2008). Intriguingly, the Rsp5 ubiquitin ligase and the Ubp3/Bre5 ubiquitin protease are involved in this preferential ribosome turnover but not bulk autophagy, supporting the existence of ribophagy (Kraft and Peter, 2008). The Ubp3–Bre5 complex interacts with the AAA ATPase Cdc48 and the ubiquitin-binding Cdc48 adaptor Ufd3 that are also required for ribophagy (Ossareh-Nazari et al., 2010). Recently, the E3 ubiquitin ligase Ltn1 has been suggested to negatively regulate ribophagy through ubiquitinating Rpl25, a 60S ribosomal subunit protein, which is also de-ubiquitinated by Ubp3 in an antagonistic action (Ossareh-Nazari et al., 2014). Whether ribosomes are recognized as disposable cargoes via ubiquitin or unknown receptor(s), and how de-ubiquitination regulates ribophagy remain to be addressed.During starvation-induced macroautophagy in mammalian cells, the timing of ribosomal degradation is different from those of other proteins and organelles, implying that bulk autophagy can even be intimately regulated in terms of cargo recognition and sequential activation (Kristensen et al., 2008).

Perspectives

Herein, we have highlighted recent progress in our understanding of organelle-specific autophagy pathways. Despite the diversity of their degradation cues and tags, the basic principles underlying organellophagy are similar among different organelles, and are likely to be universal in almost all eukaryotes. However, many of the landmark molecules for recruiting core Atg proteins are still missing, and the molecular details of organellophagy induction and termination are largely unknown.The origin of autophagosomal membranes is a fundamental, ongoing issue for all autophagy-related processes in unicellular and multicellular eukaryotes (Lamb et al., 2013). Recent imaging studies reveal the ER–mitochondria contacts as autophagosome formation sites for autophagy in mammals (Hamasaki et al., 2013) and mitophagy in yeast (Böckler and Westermann, 2014), whereas others implicate ER exit sites and the ER–Golgi intermediate compartment involving COPII vesicles for autophagy in yeast and mammals (Ge et al., 2013; Graef et al., 2013; Suzuki et al., 2013). In COS-7 cells under starvation conditions, COPII vesicles seem to localize at ER–mitochondria contacts (Tan et al., 2013), raising the possibility that these autophagosome formation sites may not be mutually exclusive. Whether degradation of other organelles utilizes the aforementioned sites for formation of autophagosomes remains to be addressed.Finally, the challenging attempts will be to decipher whether there is a cross talk between organelle biogenesis and degradation, and how the organellar quality and quantity control pathways regulate higher-order functions such as cell differentiation and development in multicellular organisms. Definitely, more stimulating discoveries are yet to come.  相似文献   

11.
Kv5, Kv6, Kv8, and Kv9 subunits: No simple silent bystanders     
Elke Bocksteins 《The Journal of general physiology》2016,147(2):105-125
  相似文献   

12.
The cell biology of disease: Lysosomal storage disorders: The cellular impact of lysosomal dysfunction     
Frances M. Platt  Barry Boland  Aarnoud C. van der Spoel 《The Journal of cell biology》2012,199(5):723-734
  相似文献   

13.
Evolution and Function of the Plant Cell Wall Synthesis-Related Glycosyltransferase Family 8     
Yanbin Yin  Huiling Chen  Michael G. Hahn  Debra Mohnen  Ying Xu 《Plant physiology》2010,153(4):1729-1746
Carbohydrate-active enzyme glycosyltransferase family 8 (GT8) includes the plant galacturonosyltransferase1-related gene family of proven and putative α-galacturonosyltransferase (GAUT) and GAUT-like (GATL) genes. We computationally identified and investigated this family in 15 fully sequenced plant and green algal genomes and in the National Center for Biotechnology Information nonredundant protein database to determine the phylogenetic relatedness of the GAUTs and GATLs to other GT8 family members. The GT8 proteins fall into three well-delineated major classes. In addition to GAUTs and GATLs, known or predicted to be involved in plant cell wall biosynthesis, class I also includes a lower plant-specific GAUT and GATL-related (GATR) subfamily, two metazoan subfamilies, and proteins from other eukaryotes and cyanobacteria. Class II includes galactinol synthases and plant glycogenin-like starch initiation proteins that are not known to be directly involved in cell wall synthesis, as well as proteins from fungi, metazoans, viruses, and bacteria. Class III consists almost entirely of bacterial proteins that are lipooligo/polysaccharide α-galactosyltransferases and α-glucosyltransferases. Sequence motifs conserved across all GT8 subfamilies and those specific to plant cell wall-related GT8 subfamilies were identified and mapped onto a predicted GAUT1 protein structure. The tertiary structure prediction identified sequence motifs likely to represent key amino acids involved in catalysis, substrate binding, protein-protein interactions, and structural elements required for GAUT1 function. The results show that the GAUTs, GATLs, and GATRs have a different evolutionary origin than other plant GT8 genes, were likely acquired from an ancient cyanobacterium (Synechococcus) progenitor, and separate into unique subclades that may indicate functional specialization.Plant cell walls are composed of three principal types of polysaccharides: cellulose, hemicellulose, and pectin. Studying the biosynthesis and degradation of these biopolymers is important because cell walls have multiple roles in plants, including providing structural support to cells and defense against pathogens, serving as cell-specific developmental and differentiation markers, and mediating or facilitating cell-cell communication. In addition to their important roles within plants, cell walls also have many economic uses in human and animal nutrition and as sources of natural textile fibers, paper and wood products, and components of fine chemicals and medicinal products. The study of the biosynthesis and biodegradation of plant cell walls has become even more significant because cell walls are the major components of biomass (Mohnen et al., 2008), which is the most promising renewable source for the production of biofuels and biomaterials (Ragauskas et al., 2006; Pauly and Keegstra, 2008). Analyses of fully sequenced plant genomes have revealed that they encode hundreds or even thousands of carbohydrate-active enzymes (CAZy; Henrissat et al., 2001; Yokoyama and Nishitani, 2004; Geisler-Lee et al., 2006). Most of these CAZy enzymes (Cantarel et al., 2009) are glycosyltransferases (GTs) or glycoside hydrolases, which are key players in plant cell wall biosynthesis and modification (Cosgrove, 2005).The CAZy database is classified into 290 protein families (www.cazy.org; release of September 2008), of which 92 are GT families (Cantarel et al., 2009). A number of the GT families have been previously characterized to be involved in plant cell wall biosynthesis. For example, the GT2 family is known to include cellulose synthases and some hemicellulose backbone synthases (Lerouxel et al., 2006), such as mannan synthases (Dhugga et al., 2004; Liepman et al., 2005), putative xyloglucan synthases (Cocuron et al., 2007), and mixed linkage glucan synthases (Burton et al., 2006). With respect to the synthesis of xylan, a type of hemicellulose, four Arabidopsis (Arabidopsis thaliana) proteins from the GT43 family, irregular xylem 9 (IRX9), IRX14, IRX9-L, and IRX14-L, and two proteins from the GT47 family, IRX10 and IRX10-L, are candidates (York and O''Neill, 2008) for glucuronoxylan backbone synthases (Brown et al., 2007, 2009; Lee et al., 2007a; Peña et al., 2007; Wu et al., 2009). In addition, three proteins have been implicated in the synthesis of an oligosaccharide thought to act either as a primer or terminator in xylan synthesis (Peña et al., 2007): two from the GT8 family (IRX8/GAUT12 [Persson et al., 2007] and PARVUS/GATL1 [Brown et al., 2007; Lee et al., 2007b]) and one from the GT47 family (FRA8/IRX7 [Zhong et al., 2005]).The GT families involved in the biosynthesis of pectins have been relatively less studied until recently. In 2006, a gene in CAZy family GT8 was shown to encode a functional homogalacturonan α-galacturonosyltransferase, GAUT1 (Sterling et al., 2006). GAUT1 belongs to a 25-member gene family in Arabidopsis, the GAUT1-related gene family, that includes two distinct but closely related families, the galacturonosyltransferase (GAUT) genes and the galacturonosyltransferase-like (GATL) genes (Sterling et al., 2006). Another GAUT gene, GAUT8/QUA1, has been suggested to be involved in pectin and/or xylan synthesis, based on the phenotypes of plant lines carrying mutations in this gene (Bouton et al., 2002; Orfila et al., 2005). It has further been suggested that multiple members of the GT8 family are galacturonosyltransferases involved in pectin and/or xylan biosynthesis (Mohnen, 2008; Caffall and Mohnen, 2009; Caffall et al., 2009).Aside from the 25 GAUT and GATL genes, Arabidopsis has 16 other family GT8 genes, according to the CAZy database, which do not seem to have the conserved sequence motifs found in GAUTs and GATLs: HxxGxxKPW and GLG (Sterling et al., 2006). Eight of these 16 genes are annotated as galactinol synthase (GolS) by The Arabidopsis Information Resource (TAIR; www.arabidopsis.org), and three of these AtGolS enzymes have been implicated in the synthesis of raffinose family oligosaccharides that are associated with stress tolerance (Taji et al., 2002). The other eight Arabidopsis GT8 genes are annotated as plant glycogenin-like starch initiation proteins (PGSIPs) in TAIR. PGSIPs have been proposed to be involved in the synthesis of primers necessary for starch biosynthesis (Chatterjee et al., 2005). Hence, the GT8 family is a protein family consisting of enzymes with very distinct proven and proposed functions. Indeed, a suggestion has been made to split the GT8 family into two groups (Sterling et al., 2006), namely, the cell wall biosynthesis-related genes (GAUTs and GATLs) and the non-cell wall synthesis-related genes (GolSs and PGSIPs).We are interested in further defining the functions of the GAUT and GATL proteins in plants, in particular their role(s) in plant cell wall synthesis. The apparent disparate functions of the GT8 family (i.e. the GAUTs and GATLs as proven and putative plant cell wall polysaccharide biosynthetic α-galacturonosyltransferases, the eukaryotic GolSs as α-galactosyltransferases that synthesize the first step in the synthesis of the oligosaccharides stachyose and raffinose, the putative PGSIPs, and the large bacterial GT8 family of diverse α-glucosyltransferases and α-galactosyltransferases involved in lipopolysaccharide and lipooligosaccharide synthesis) indicate that the GT8 family members are involved in several unique types of glycoconjugate and glycan biosynthetic processes (Yin et al., 2010). This observation led us to ask whether any of the GT8 family members are sufficiently closely related to GAUT and GATL genes to be informative regarding GAUT or GATL biosynthetic function(s) and/or mechanism(s).To investigate the relatedness of the members of the GT8 gene family, we carried out a detailed phylogenetic analysis of the entire GT8 family in 15 completely sequenced plant and green algal genomes (AbbreviationCladeSpeciesGenome PublishedDownloaded frommpcGreen algaeMicromonas pusilla CCMP1545Worden et al. (2009)JGI version 2.0mprGreen algaeMicromonas strain RCC299Worden et al. (2009)JGI version 2.0olGreen algaeOstreococcus lucimarinusPalenik et al. (2007)JGI version 1.0otGreen algaeOstreococcus tauriDerelle et al. (2006)JGI version 1.0crGreen algaeChlamydomonas reinhardtiiMerchant et al. (2007)JGI version 3.0vcGreen algaeVolvox carteri f. nagariensisNoJGI version 1.0ppMossPhyscomitrella patens ssp. patensRensing et al. (2008)JGI version 1.1smSpike mossSelaginella moellendorffiiNoJGI version 1.0ptDicotPopulus trichocarpaTuskan et al. (2006)JGI version 1.1atDicotArabidopsis thalianaArabidopsis Genome Initiative (2000)TAIR version 9.0vvDicotVitis viniferaJaillon et al. (2007)http://www.genoscope.cns.fr/gmDicotGlycine maxSchmutz et al. (2010)JGI version 1.0osMonocotOryza sativaGoff et al. (2002); Yu et al. (2002)TIGR version 6.1sbMonocotSorghum bicolorPaterson et al. (2009)JGI version 1.0bdMonocotBrachypodium distachyonVogel et al. (2010)JGI version 1.0Open in a separate window  相似文献   

14.
TANG1, Encoding a Symplekin_C Domain-Contained Protein,Influences Sugar Responses in Arabidopsis     
Leiying Zheng  Li Shang  Xing Chen  Limin Zhang  Yan Xia  Caroline Smith  Michael W. Bevan  Yunhai Li  Hai-Chun Jing 《Plant physiology》2015,168(3):1000-1012
  相似文献   

15.
The Golgi and the centrosome: building a functional partnership     
Christine Sütterlin  Antonino Colanzi 《The Journal of cell biology》2010,188(5):621-628
The mammalian Golgi apparatus is characterized by a ribbon-like organization adjacent to the centrosome during interphase and extensive fragmentation and dispersal away from the centrosome during mitosis. It is not clear whether this dynamic association between the Golgi and centrosome is of functional significance. We discuss recent findings indicating that the Golgi–centrosome relationship may be important for directional protein transport and centrosome positioning, which are both required for cell polarization. We also summarize our current knowledge of the link between Golgi organization and cell cycle progression.

Introduction

The Golgi apparatus plays a central role in the secretory pathway. Newly synthesized proteins are transported from the ER to the Golgi, where they are posttranslationally modified. They are sorted into carriers for delivery to the plasma membrane or the endosomal–lysosomal system. The basic structural unit of the Golgi apparatus is a stack of flattened cisternae that is morphologically conserved among most species. In mammalian cells, individual Golgi stacks are connected laterally to form a continuous membranous system called the Golgi ribbon, which is located in close physical proximity to the centrosome (Fig. 1, left).Open in a separate windowFigure 1.The spatial relationship between the Golgi and the centrosome during the mammalian cell cycle. Golgi (red, stained with antibodies to GM130) and centrosome (green, stained with antibodies to centrin) staining of nonsynchronized bone cancer cells (U2-OS) shows the physical proximity between these two organelles during interphase (left) and its temporary loss during mitosis (right). Bar, 10 µm.The centrosome functions as the major microtubule-organizing center of the cell and plays an important role in cell polarization and ciliogenesis (Bettencourt-Dias and Glover, 2007). In a newly formed daughter cell, this nonmembrane-bound organelle is composed of a pair of centrioles that is surrounded by a cloud of electron-dense material called the pericentriolar matrix. γ-Tubulin ring complexes (γ-TuRCs) in the pericentriolar matrix allow the centrosome to nucleate the radial array of interphase microtubules whose minus ends are embedded in the centrosome and whose plus ends extend toward the cell periphery. After centrosome duplication in S phase, the two centrosomes move to opposite poles of the cell and become the spindle poles from which spindle microtubules grow. Centrosomes are generally located in the cell center close to the nucleus, although this central position is lost in response to a polarization stimulus, which prompts centrosomes to reorient toward the leading edge of the cell (Pouthas et al., 2008). In most cell types, centrosome reorientation is critical for the ability of cells to polarize and migrate (Yvon et al., 2002). The centrosome is also linked to ciliogenesis because one of its centrioles is converted into the basal body from which a primary cilium extends (D''Angelo and Franco, 2009).The spatial relationship between the Golgi apparatus and the centrosome is altered by changes in Golgi organization that occur during the cell cycle (Fig. 1). These two organelles are only adjacent in interphase when the Golgi apparatus is arranged as a ribbon in the pericentriolar region (Colanzi et al., 2003). In contrast, Golgi membranes are fragmented and dispersed throughout the cytoplasm during mitosis. Intriguingly, the pericentriolar localization of the Golgi is a feature typical of some eukaryotic cells, ranging from mammalian and amphibian cells (Thyberg and Moskalewski, 1999; Reilein et al., 2003) to amoeba (Rehberg et al., 2005). However, other eukaryotes, including plants and flies, have isolated Golgi stacks (Stanley et al., 1997; Nebenführ and Staehelin, 2001) or isolated cisternae in the case of Saccharomyces cerevisiae that are scattered throughout the cytoplasm without an obvious connection with the centrosome (Preuss et al., 1992).In this paper, we review recent findings indicating that the relationship between the Golgi and the centrosome in interphase is important for cell polarization. We also summarize the current understanding of how Golgi–centrosome interactions during mitosis affect cell division.

Are there functional interactions between the Golgi and the centrosome during interphase?

Golgi membranes are actively positioned in the pericentriolar position.

The localization of the mammalian Golgi ribbon next to the centrosome requires the microtubule and actin cytoskeleton (Brownhill et al., 2009). Microtubules have a dual role in organizing the pericentriolar Golgi ribbon. First, the subset of microtubules that is nucleated at the Golgi is necessary for the assembly of Golgi fragments into a connected ribbon in the cell periphery (Miller et al., 2009). Second, centrosomal microtubules provide the tracks along which Golgi membranes are transported to the cell center (Cole et al., 1996). Both steps depend on the minus end–directed motor complex dynein (Burkhardt et al., 1997; Miller et al., 2009). The actin cytoskeleton is also involved in localizing Golgi membranes. Actin fibers, which have been detected at the Golgi complex, are required for the maintenance of the pericentriolar position of this organelle by providing tracks for actin-based motors (Valderrama et al., 1998; Sahlender et al., 2005; Vicente-Manzanares et al., 2007). Actin fibers and microtubules are coordinated by proteins that associate with both cytoskeletal elements such as WHAMM, a Golgi-bound actin-nucleating factor, and MACF1, a microtubule–actin cross-linking protein (Lin et al., 2005; Campellone et al., 2008).Golgi organization and localization in the pericentriolar region also depend on Golgi-associated proteins (Ramirez and Lowe, 2009). Their depletion produces defects in Golgi organization ranging from a disconnected Golgi ribbon in the pericentriolar region (Puthenveedu et al., 2006) to dispersed ministacks in the cytoplasm (Diao et al., 2003; Yadav et al., 2009).

Table I.

Golgi-associated proteins that control the pericentriolar position of the Golgi apparatus
Golgi-associated proteinsReference
Structural Golgi proteins
Cog3Zolov and Lupashin, 2005
GCC185Derby et al., 2007
GCP60 (ACBD3)Sohda et al., 2001
GM130Marra et al., 2007
Golgin-45Short et al., 2001
Golgin-84Diao et al., 2003
Golgin-97Lu et al., 2004
Golgin-160Yadav et al., 2009
Golgin-245Yoshino et al., 2005
GRASP55Feinstein and Linstedt, 2008
GRASP65Puthenveedu et al., 2006
p115Sohda et al., 2005
Membrane traffic
RINT-1Sun et al., 2007
Syntaxin 5Suga et al., 2005
ZW10Sun et al., 2007
Cytoskeleton regulators and motors
ARHGAP10Dubois et al., 2005
CG-NAP/AKAP450Takahashi et al., 1999
CLASP2Efimov et al., 2007
Coronin 7Rybakin et al., 2006
GMAP-210Ríos et al., 2004
FTCDGao and Sztul, 2001
Hook3Walenta et al., 2001
MACF1bLin et al., 2005
Myosin IIVicente-Manzanares et al., 2007
Myosin VISahlender et al., 2005
OptineurinSahlender et al., 2005
p50/dynamitinRoghi and Allan, 1999
WHAMMCampellone et al., 2008
Kinases and enzymes
Cdk5 (kinase)Sun et al., 2008
ORP9 (lipid transfer)Ngo and Ridgway, 2009
PKA (kinase)Bejarano et al., 2006
PKD1 (kinase)Díaz Añel and Malhotra, 2005
Sac1 (PI phosphatase)Liu et al., 2008
Open in a separate windowAlthough the position of the Golgi next to the centrosome is actively maintained, it does not appear to be critical for basic Golgi functions. For example, membrane trafficking and the modification of secretory proteins are unaffected when the Golgi ribbon is severed into individual ministacks (Cole et al., 1996; Diao et al., 2003; Yadav et al., 2009). Furthermore, organisms such as S. cerevisiae secrete proteins with high efficiency, although their Golgi membranes are never pericentriolar (Preuss et al., 1992). Thus, the physiological role of the pericentrosomal positioning of the mammalian Golgi apparatus remains a major unanswered question.

An emerging role for Golgi–centrosome association in polarized secretion.

New studies indicate that the relationship between the Golgi and the centrosome may be important for specialized functions of mammalian cells. A prominent example is cell polarization, which is a prerequisite for cell migration (Li et al., 2005). Cell polarization depends on directional protein transport along Golgi-nucleated microtubules as well as centrosome reorientation toward the leading edge of the cell, which both appear to be affected by interactions between the Golgi and the centrosome.In a recent study, Yadav et al. (2009) investigated the role of the pericentriolar Golgi ribbon in directional transport and cell polarization. Depletion of each of the two structural proteins of the golgin family, GMAP210 and Golgin-160, disrupted the ribbon-like structure of the Golgi and led to isolated ministacks in the cytoplasm. These dispersed stacks were competent of general protein transport to the cell surface. However, there were defects in directional protein transport, as shown by the failure to secrete vesicular stomatitis virus G protein in a directional manner toward the leading edge of a cell and the inability of these cells to migrate in a wound-healing assay. These results indicate that the pericentriolar Golgi ribbon is critical for directional protein transport, although it is not clear whether it is the ribbon-like organization or the position next to the centrosome that is important.Golgi–centrosome interactions may also contribute to cell polarization through regulatory effects on centrosome positioning. Both the centrosome and the Golgi apparatus undergo reorientation toward the leading edge of a stimulated cell. Bisel et al. (2008) found that centrosome reorientation depends on phosphorylation of the Golgi protein GRASP65, which is proposed to promote Golgi stack disassembly (Wang et al., 2003; Yoshimura et al., 2005). In this study, expression of nonphosphorylatable forms of GRASP65 prevented Golgi and centrosome reorientation toward the leading edge and cell migration. Intriguingly, this block was overcome when Golgi membranes were artificially fragmented, indicating that Golgi membranes have to be remodeled to allow the coordinated reorientation of the centrosome and the Golgi. Thus, the ability of the Golgi to reorganize affects the positioning of the centrosome (Bisel et al., 2008).The peripheral Golgi protein, GM130, is an additional critical factor in the regulation of cell polarization (Preisinger et al., 2004; Kodani et al., 2009; Rivero et al., 2009). There are at least four reasons to explain why depletion of GM130 prevents cells from polarizing and migrating in wound-healing assays (Kodani et al., 2009). First, Kodani and Sütterlin (2008) showed that GM130 depletion altered the organization of the centrosome so that it was no longer able to nucleate microtubules or to reorient in response to a polarization stimulus. Second, GM130-dependent centrosome regulation involved the small GTPase Cdc42 (Kodani et al., 2009), a known regulator of cell polarization (Etienne-Manneville, 2006; Kodani et al., 2009). Third, Rivero et al. (2009) identified a novel role for GM130 in microtubule nucleation at the Golgi, which required GM130-dependent recruitment of the microtubule nucleation factor AKAP450 to the Golgi (Rivero et al., 2009). Golgi-nucleated microtubules, which were first identified in in vitro studies (Chabin-Brion et al., 2001), are preferentially oriented toward the leading edge of a motile cell and are necessary for directional protein transport (Fig. 2; Rivero et al., 2009). Fourth, GM130 binds and activates the protein kinase YSK1, which has a known role in cell migration (Preisinger et al., 2004). Thus, GM130 may affect cell polarization and migration through effects on centrosome organization, Cdc42 activation, microtubule nucleation at the Golgi, and YSK1 activation.Open in a separate windowFigure 2.Golgi- and centrosome-nucleated microtubules in cell migration. The centrosome nucleates a radial array of microtubules (red) whose minus ends (−) are anchored at the centrosome and whose plus ends (+) extend into the cell periphery. This population of microtubules depends on γ-TuRC complexes and the large scaffold protein AKAP450 for their nucleation and functions in maintaining the pericentriolar localization of the Golgi ribbon by a dynein-mediated mechanism (closed arrows). In contrast, the Golgi apparatus nucleates microtubules (brown) that extend asymmetrically toward the leading edge of a migrating cell. Microtubule nucleation at the Golgi requires the peripheral Golgi protein GM130, which recruits AKAP450 and γ-TuRC complexes to the Golgi apparatus. Golgi-nucleated microtubules are coated with CLASP proteins and are necessary for the formation of the Golgi ribbon from dispersed stacks. In addition, they are required for cell migration by facilitating polarized protein transport to the leading edge of a cell (open arrows).The formation of a primary cilium is another process that involves interactions between the Golgi and the centrosome. During ciliogenesis, the centrosome moves to the plasma membrane, where one of its centrioles becomes the basal body from which the primary cilium extends. IFT20, a critical component of the intraflagellar transport machinery that is required for formation and extension of the cilium (Follit et al., 2006), localizes to the Golgi by binding to the structural Golgi protein GMAP210. Loss of either IFT20 or GMAP210 impairs ciliogenesis (Follit et al., 2006, 2008), which supports a role for Golgi-localized IFT20 in protein sorting at the Golgi to produce transport carriers involved in the formation of a primary cilium. A similar role in directing specific cargo molecules to the ciliary membrane has been proposed for the small GTPase Rab8, which also localizes to the Golgi and the basal body (Nachury et al., 2007). Collectively, these new findings are intriguing, as they provide support for a functional link between the Golgi and the centrosome.

Are there functional interactions between the Golgi and the centrosome during mitosis?

Regulation of mitotic Golgi reorganization from the centrosome.

The physical proximity of the Golgi apparatus and the centrosome is transiently lost during mitosis when Golgi membranes undergo extensive fragmentation. This dramatic change in Golgi structure is concomitant with a block in secretory trafficking and the reorganization of the microtubule cytoskeleton (Colanzi et al., 2003). Although the Golgi and the centrosome are physically separate at this stage of the cell cycle, there is evidence for functional interactions between these two organelles, which may control progression through mitosis.Many studies have identified possible links between mitotic Golgi fragmentation and the centrosome. For instance, breaking the Golgi ribbon into its constituent stacks during G2 requires the activity of the protein kinase Plk3 (Xie et al., 2004; López-Sánchez et al., 2009), which localizes to the centrosome and spindle poles (Xie et al., 2004; Jiang et al., 2006). The subsequent conversion of Golgi stacks into small, highly dispersed fragments (Jesch et al., 2001; Altan-Bonnet et al., 2006) and vesicular/tubular clusters next to astral spindle microtubules (Shima et al., 1998; Wei and Seemann, 2009) is regulated by two mitotic kinases, Cdk1 and Plk1, which are both associated with the centrosome (Fig. 3; Bailly et al., 1989; Dai and Cogswell, 2003). These findings suggest that components of the centrosome, spindle poles, or the spindle may initiate a signaling pathway that leads to the fragmentation of the Golgi and that may help coordinate Golgi dynamics with cell cycle progression. However, these regulatory factors also exist in the cytosol, and possible roles of cytosolic pools of Cdk1 and Plk3 in mitotic Golgi fragmentation have not been excluded.Open in a separate windowFigure 3.Golgi fragmentation during mitosis. The mammalian Golgi apparatus (green) forms an interconnected ribbon adjacent to the centrosome (red) and the nucleus (blue). It nucleates a population of microtubules that is necessary for polarized protein transport. Plus (+) and minus ends (−) are indicated. The activities of the protein kinases Plk3 and MEK1 and the fission protein BARS are required to convert the ribbon structure into isolated stacks in late G2 and prophase. In metaphase, the isolated stacks are further fragmented by a Plk1- and Cdk1-dependent mechanism, producing vesicular/tubular membranes that are dispersed throughout the cytoplasm. During this process, ribbon determinants, which are proteins required for postmitotic Golgi ribbon formation, remain associated with the mitotic spindle for their partitioning into daughter cells. Centrosome-associated regulators of mitotic Golgi fragmentation are labeled in red. Regulators of Golgi fragmentation that are not associated with the centrosome are labeled in black.Further support for functional interactions between the Golgi and the centrosome during mitosis stems from a novel study on spindle-dependent reassembly of the Golgi ribbon after mitosis (Wei and Seemann, 2009). In a series of elegant experiments, Wei and Seemann (2009) demonstrated that the spindle is required for the postmitotic reformation of the Golgi ribbon. They induced asymmetric cell division so that the entire spindle segregated into only one daughter cell. Although Golgi membranes assembled into stacks in both daughter cells, they only formed a ribbon in the cell that inherited the spindle. Ribbon formation in the spindle-free cell required coinjection of Golgi extracts and tubulin or the addition of spindle-containing fractions. Collectively, these results suggest that Golgi ribbon formation occurs in two steps, with the initial assembly into stacks being mediated by factors that are partitioned by a spindle-independent mechanism. The subsequent formation of the Golgi ribbon from individual stacks, however, has an additional requirement for ribbon determinants, which are likely to be Golgi-associated proteins inherited with the spindle. Possible candidates include regulators of Golgi dynamics and the secretory pathway that have been identified in preparations of the spindle matrix (Ma et al., 2009).

Significance of the loss of Golgi–centrosome proximity during mitosis.

Several studies have identified an unexpected link between mitotic Golgi fragmentation and cell cycle progression (Sütterlin et al., 2002; Hidalgo Carcedo et al., 2004; Preisinger et al., 2005). For example, interfering with mitotic Golgi disassembly by blocking the function of the peripheral Golgi protein GRASP65 or the fission protein BARS resulted in cell cycle arrest in G2 (Sütterlin et al., 2002; Hidalgo Carcedo et al., 2004). Intriguingly, breaking the ribbon into isolated stacks, which occurs in G2, is sufficient to overcome this cell cycle arrest and allows cells to enter mitosis (Colanzi et al., 2007; Feinstein and Linstedt, 2007). It is not known how and why the presence of an intact pericentriolar Golgi ribbon prevents mitotic entry. The existence of a Golgi checkpoint, which monitors the correct inheritance of the Golgi complex, has been proposed because these inhibitory effects are not caused by activation of the DNA damage checkpoint (Sütterlin et al., 2002; Hidalgo Carcedo et al., 2004). It is conceivable that severing the Golgi ribbon in G2 separates ribbon determinants from the rest of the Golgi so that they can cosegregate with the spindle (Fig. 3). Such a mechanism for spindle-dependent Golgi inheritance would ensure that both daughter cells inherit the ability to form a Golgi ribbon and, thus, to transport proteins in a polarized manner. By analogy to the spindle checkpoint, which controls the exit from mitosis by monitoring the correct binding of spindle microtubules to kinetochores, this Golgi checkpoint may assess binding of spindle microtubules to these putative ribbon determinants to regulate entry into mitosis.

Golgi-dependent regulation of the spindle and mitotic progression.

A series of recent studies has identified a requirement for specific Golgi-associated proteins in the formation of a bipolar spindle (Chang et al., 2005), the putative Golgi stacking factor, GRASP65 (Sütterlin et al., 2005), a regulator of the spindle checkpoint, RINT-1 (Lin et al., 2007), and the phosphatidylinositide phosphatase, Sac1 (Burakov et al., 2008). Depletion of any one of these proteins leads to multipolar spindles and mitotic cell death. For example, RNAi-mediated knockdown of Sac1 resulted in disorganization of the Golgi apparatus and mitotic defects characterized by multiple mechanically active spindles (Liu et al., 2008). Similarly, loss of GRASP65 led to the formation of multipolar spindles and mitotic arrest followed by cell death (Sütterlin et al., 2005). The molecular mechanisms by which Golgi-associated proteins regulate spindle formation are not known. Also, it is not known whether Golgi components control spindle formation when Golgi membranes are in the form of a pericentriolar ribbon, isolated stacks, or small fragments.

Table II.

Golgi-associated proteins with a role in regulating centrosome and spindle function
ProteinFunctionDepletion phenotypeReference
Sac1Lipid phosphataseMultiple mechanically active spindlesLiu et al., 2008
GM130GolginAberrant centrosome, multipolar spindlesKodani and Sütterlin, 2008
GRASP65Golgi matrixMultipolar spindles, mitotic cell deathSütterlin et al., 2005
RINT-1Membrane trafficMultipolar spindles, mitotic cell deathSun et al., 2007
Tankyrase-1ADP ribosyl transferaseMultipolar spindles, mitotic cell deathChang et al., 2005
Rab6′GTPaseMetaphase block, SAC activationMiserey-Lenkei et al., 2006
ClathrinVesicle coatDefects in chromosome congression, SAC activationRoyle et al., 2005
Open in a separate windowSAC, spindle assembly checkpoint.In addition to Golgi-dependent effects on spindle formation, other mitotic events are also regulated by disassembly of Golgi stacks during prophase and prometaphase. Indeed, this disassembly step correlates with the release of several peripheral proteins from Golgi membranes to carry out specific functions during mitosis. For instance, clathrin dissociates from the Golgi complex and from endocytic vesicles during mitosis and localizes to the spindle pole where it stabilizes mitotic spindle fibers involved in chromosome segregation (Royle et al., 2005). The small GTPase, Rab6A′, is also released from the Golgi during mitotic Golgi fragmentation (Miserey-Lenkei et al., 2006). If this dynamic behavior of Rab6A′ is inhibited, cells are no longer able to progress through mitosis and are blocked in metaphase through activation of the spindle checkpoint. Another example is the Golgi-associated protein ACBD3, whose release and cytoplasmic dispersal during mitotic Golgi breakdown is necessary for the activation of Numb in the regulation of asymmetric cell division (Zhou et al., 2007). Thus, in addition to facilitating the partitioning of Golgi membranes into the daughter cells, Golgi fragmentation may provide a unique mechanism for the regulation of signaling pathways that involve Golgi-associated components. In the case of ACBD3 and Rab6A′, Golgi fragmentation may relieve inhibitory effects that are either the result of proximity with the centrosome or the organization of the Golgi ribbon.

Conclusions

There is increasing evidence that the relationship between the Golgi apparatus and the centrosome in mammalian cells extends beyond physical proximity and involves functional interactions. Several features of this Golgi–centrosome relationship can be surmised from the recent studies reviewed. This relationship appears to be bidirectional because components of each organelle are able to influence the function of the other organelle. For example, Golgi proteins are necessary for centrosome organization and positioning (Chang et al., 2005; Sütterlin et al., 2005; Kodani and Sütterlin, 2008), whereas centrosome-nucleated microtubules are required for pericentriolar Golgi positioning (Corthésy-Theulaz et al., 1992; Cole et al., 1996). Importantly, these functional interactions affect fundamental cellular processes such as cell polarization and progression through mitosis (Sütterlin et al., 2002; Yadav et al., 2009). Intriguingly, there is evidence for functional interactions when the Golgi and the centrosome are in physical proximity during interphase but also during mitosis when they are physically separate.What is the functional significance of the physical proximity between the Golgi and the centrosome? One possibility is that it may enhance the efficiency of signaling between the Golgi and centrosome and thereby facilitate directional protein transport. The Golgi apparatus is well known for its role in the exocytic pathway, and Golgi membranes, the intermediate compartment, and late endosomes are concentrated in the centrosomal area in mammalian cells (Marie et al., 2009). Thus, the centrosomal area may serve as a traffic hub, allowing integrated regulation of exocytic and endocytic transport routes for polarized delivery of cargo. In support of this idea, species in which Golgi membranes are not adjacent to the centrosome use alternative strategies for transporting proteins in a directional manner. For example, polarized secretion in Drosophila melanogaster is achieved by targeting mRNA to specific transitional ER–Golgi units in which the cargo is synthesized and secreted locally (Herpers and Rabouille, 2004).Why has it taken so long to reveal functional Golgi–centrosome interactions during cell division? The phenomenon of a pericentriolar interphase Golgi ribbon, which is fragmented and dispersed during mitosis, is mainly seen in mammalian cells. Therefore, the significance of this dynamic spatial relationship cannot be studied in a more genetically tractable system such as yeast or Drosophila in which genome-wide screens can be readily performed. Furthermore, there has been a lack of tools to separate the Golgi and centrosome without affecting the functions of these organelles. Some recent studies have used new approaches such as severing the Golgi ribbon by depleting structural golgins, but there are still experimental limitations. For example, an intact Golgi ribbon cannot simply be displaced from the pericentriolar region, which makes it difficult to directly test the significance of Golgi localization versus organization. In addition, Golgi fragmentation, as induced by the depletion of structural Golgi proteins, is a multifactorial process that is marked by both the loss of the Golgi ribbon and dispersal from the pericentriolar position. The limited availability of experimental tools makes it difficult to separate these processes, which has hampered efforts to dissect their individual contributions to the Golgi–centrosome partnership. Also, until a recent study (Kodani et al., 2009), a molecular pathway linking the Golgi and the centrosome during interphase had not been described. For these reasons, it has been difficult to experimentally alter Golgi–centrosome proximity and assay the effects.Although progress has been made, there are many unresolved questions about the Golgi–centrosome relationship during the cell cycle. For example, is there a single bidirectional regulatory pathway between the Golgi and the centrosome, or are there separate signaling pathways in each direction? Are there differences in signaling between these organelles during interphase when the organelles are adjacent and in mitosis when they are physically separate? There are also more specific unanswered questions. For example, how do Golgi proteins control spindle formation? Which factors on the mitotic spindle regulate postmitotic reassembly of the Golgi? How does the organization of the Golgi apparatus control progression through the cell cycle? Is there a Golgi organization checkpoint, and what does it monitor? The answers to these questions will help us better understand the significance of Golgi–centrosome interactions and could lead to the development of novel approaches for the treatment of several important diseases, including cancer.  相似文献   

16.
Calcineurin B-Like Protein-Interacting Protein Kinase CIPK21 Regulates Osmotic and Salt Stress Responses in Arabidopsis     
Girdhar K. Pandey  Poonam Kanwar  Amarjeet Singh  Leonie Steinhorst  Amita Pandey  Akhlilesh K. Yadav  Indu Tokas  Sibaji K. Sanyal  Beom-Gi Kim  Sung-Chul Lee  Yong-Hwa Cheong  J?rg Kudla  Sheng Luan 《Plant physiology》2015,169(1):780-792
The role of calcium-mediated signaling has been extensively studied in plant responses to abiotic stress signals. Calcineurin B-like proteins (CBLs) and CBL-interacting protein kinases (CIPKs) constitute a complex signaling network acting in diverse plant stress responses. Osmotic stress imposed by soil salinity and drought is a major abiotic stress that impedes plant growth and development and involves calcium-signaling processes. In this study, we report the functional analysis of CIPK21, an Arabidopsis (Arabidopsis thaliana) CBL-interacting protein kinase, ubiquitously expressed in plant tissues and up-regulated under multiple abiotic stress conditions. The growth of a loss-of-function mutant of CIPK21, cipk21, was hypersensitive to high salt and osmotic stress conditions. The calcium sensors CBL2 and CBL3 were found to physically interact with CIPK21 and target this kinase to the tonoplast. Moreover, preferential localization of CIPK21 to the tonoplast was detected under salt stress condition when coexpressed with CBL2 or CBL3. These findings suggest that CIPK21 mediates responses to salt stress condition in Arabidopsis, at least in part, by regulating ion and water homeostasis across the vacuolar membranes.Drought and salinity cause osmotic stress in plants and severely affect crop productivity throughout the world. Plants respond to osmotic stress by changing a number of cellular processes (Xiong et al., 1999; Xiong and Zhu, 2002; Bartels and Sunkar, 2005; Boudsocq and Lauriére, 2005). Some of these changes include activation of stress-responsive genes, regulation of membrane transport at both plasma membrane (PM) and vacuolar membrane (tonoplast) to maintain water and ionic homeostasis, and metabolic changes to produce compatible osmolytes such as Pro (Stewart and Lee, 1974; Krasensky and Jonak, 2012). It has been well established that a specific calcium (Ca2+) signature is generated in response to a particular environmental stimulus (Trewavas and Malhó, 1998; Scrase-Field and Knight, 2003; Luan, 2009; Kudla et al., 2010). The Ca2+ changes are primarily perceived by several Ca2+ sensors such as calmodulin (Reddy, 2001; Luan et al., 2002), Ca2+-dependent protein kinases (Harper and Harmon, 2005), calcineurin B-like proteins (CBLs; Luan et al., 2002; Batistič and Kudla, 2004; Pandey, 2008; Luan, 2009; Sanyal et al., 2015), and other Ca2+-binding proteins (Reddy, 2001; Shao et al., 2008) to initiate various cellular responses.Plant CBL-type Ca2+ sensors interact with and activate CBL-interacting protein kinases (CIPKs) that phosphorylate downstream components to transduce Ca2+ signals (Liu et al., 2000; Luan et al., 2002; Batistič and Kudla, 2004; Luan, 2009). In several plant species, multiple members have been identified in the CBL and CIPK family (Luan et al., 2002; Kolukisaoglu et al., 2004; Pandey, 2008; Batistič and Kudla, 2009; Weinl and Kudla, 2009; Pandey et al., 2014). Involvement of specific CBL-CIPK pair to decode a particular type of signal entails the alternative and selective complex formation leading to stimulus-response coupling (D’Angelo et al., 2006; Batistič et al., 2010).Several CBL and CIPK family members have been implicated in plant responses to drought, salinity, and osmotic stress based on genetic analysis of Arabidopsis (Arabidopsis thaliana) mutants (Zhu, 2002; Cheong et al., 2003, 2007; Kim et al., 2003; Pandey et al., 2004, 2008; D’Angelo et al., 2006; Qin et al., 2008; Tripathi et al., 2009; Held et al., 2011; Tang et al., 2012; Drerup et al., 2013; Eckert et al., 2014). A few CIPKs have also been functionally characterized by gain-of-function approach in crop plants such as rice (Oryza sativa), pea (Pisum sativum), and maize (Zea mays) and were found to be involved in osmotic stress responses (Mahajan et al., 2006; Xiang et al., 2007; Yang et al., 2008; Tripathi et al., 2009; Zhao et al., 2009; Cuéllar et al., 2010).In this report, we examined the role of the Arabidopsis CIPK21 gene in osmotic stress response by reverse genetic analysis. The loss-of-function mutant plants became hypersensitive to salt and mannitol stress conditions, suggesting that CIPK21 is involved in the regulation of osmotic stress response in Arabidopsis. These findings are further supported by an enhanced tonoplast targeting of the cytoplasmic CIPK21 through interaction with the vacuolar Ca2+ sensors CBL2 and CBL3 under salt stress condition.  相似文献   

17.
A Rice Ca2+ Binding Protein Is Required for Tapetum Function and Pollen Formation     
Pil Joong Chung  Bong Soo Park  Huan Wang  Jun Liu  In-Cheol Jang  Nam-Hai Chua 《Plant physiology》2016,172(3):1772-1786
  相似文献   

18.
SUCROSE NONFERMENTING1-RELATED PROTEIN KINASE2.6, an Ortholog of OPEN STOMATA1, Is a Negative Regulator of Strawberry Fruit Development and Ripening     
Yu Han  Ruihong Dang  Jinxi Li  Jinzhu Jiang  Ning Zhang  Meiru Jia  Lingzhi Wei  Ziqiang Li  Bingbing Li  Wensuo Jia 《Plant physiology》2015,167(3):915-930
  相似文献   

19.
Plasticity of cell migration: a multiscale tuning model     
Peter Friedl  Katarina Wolf 《The Journal of cell biology》2010,188(1):11-19
Cell migration underlies tissue formation, maintenance, and regeneration as well as pathological conditions such as cancer invasion. Structural and molecular determinants of both tissue environment and cell behavior define whether cells migrate individually (through amoeboid or mesenchymal modes) or collectively. Using a multiparameter tuning model, we describe how dimension, density, stiffness, and orientation of the extracellular matrix together with cell determinants—including cell–cell and cell–matrix adhesion, cytoskeletal polarity and stiffness, and pericellular proteolysis—interdependently control migration mode and efficiency. Motile cells integrate variable inputs to adjust interactions among themselves and with the matrix to dictate the migration mode. The tuning model provides a matrix of parameters that control cell movement as an adaptive and interconvertible process with relevance to different physiological and pathological contexts.

Introduction

Cell migration is a complex and heterogeneous process executed by all nucleated cell types at a given time window of their development. For most cells, including epithelial, stromal, and neuronal cells, migration phases are confined to morphogenesis and cease with terminal differentiation toward intact tissue to become reactivated only for tissue regeneration or neoplastic processes. For other cell types, such as leukocytes, migration is integral to their function and maintained throughout their life span. Some cell types migrate only in the context of a defined substrate, such as epithelial cells moving along a basement membrane but not through interstitial tissues, whereas other cell types, including leukocytes, are versatile, as they interact with and migrate within virtually any substrate present in the body. Thus, although the same basic process is executed (i.e., cell translocation along or through tissue structures), each cell type exerts migration in different contexts using distinct molecular repertoires and extracellular guidance cues. We here summarize extra- and intracellular molecular parameters that regulate cell migration and integrate them into a parameter “matrix” to better classify how cell migration modes are being both achieved and modulated.

The modes of cell migration

The mode of cell migration was originally classified based on the morphology of migration patterns. This terminology was then extended to include molecular parameters, such as cytoskeletal organization, the type of cell–matrix interaction and force generation, and the modification of the tissue structure imposed by migrating cells (Friedl et al., 1998b; Thiery, 2002; Friedl, 2004; Lämmermann and Sixt, 2009; Sanz-Moreno and Marshall, 2009). As main categories, cell move either individually (amoeboid or mesenchymal) or collectively (the migration of cohesive multicellular units; Fig. 1 and Friedl, 2004). Although these terms are arguably arbitrary and the molecular discrimination between the certain modes is incomplete, they help to simplify and categorize an otherwise diffuse literature and allow dissection of the molecular machineries underlying each mode.Open in a separate windowFigure 1.Cell morphologies, migration modes, and transitions. The nomenclature of interstitial migration modes is based on typical cell morphology (rounded or spindle-shaped) and pattern (individual, loosely connected, or collective). Each migration mode is governed by a set of molecular mechanisms (see details in Fig. 2), the regulation of which can change the style of migration. Most widely studied examples for alterations of migration mode are the mesenchymal-to-amoeboid transition or the collective-to-individual transition. The thick gray arrows indicate the direction of migration.

Table I.

Different migration modes and selected determinants
Migration modeCell typesECM determinantsCell determinantsRelated transitionsReferences
Single
    Amoeboid, blebbyZebrafish macrophage, some stem cellsPoorly adhesive; soft embryonic connective tissue; obligate 3DAsymmetric bleb-rich cortical actomyosin cytoskeleton, low polarity; low migration speed (below 1 µm/min)Blebby-to-pseudopodal transitionsBlaser et al., 2006; Yoshida and Soldati, 2006
    Amoeboid, pseudopodalLeukocytes, including dendritic cells; Dictyostelium discoideumLoose primordial or mature connective tissue; 2D or 3DPoorly adhesive, no formation of focal adhesions; Rac-driven anterior protrusion with counterbalance by Rho/ROCK in other cell parts; relatively rapid migration (10 µm/min)Amoeboid-to-mesenchymal transitionYoshida and Soldati, 2006; Lämmermann et al., 2008
    MesenchymalFibroblasts, neural crest cells, sarcoma cells, dedifferentiated cancer cells of different originLoose or dense primordial or mature connective tissue; usually associated with fibrin or collagen remodelingModerately to highly adhesive; focal interactions with ECM; high contractility; high anterior Rac activity counterbalanced by Rho in other cell parts; slow migration (0.1–1 µm/min)Mesenchymal-to-amoeboid transition; mesenchymal-to-epithelial/collective transitionWolf et al., 2003a, 2007; Grinnell, 2008; Paňková et al., 2009; Thiery, 2002
Multicellular
    Chain migration, cell streamingNeural crest cells, fibroblastsJoint ECM tracks?Individual cells with temporary tiplike cell-cell contactsMigration arrest and integration into terminal tissueDavis and Trinkaus, 1981; Kulesa and Fraser, 2000
    CollectiveDictyostelium at slug stage, lateral line (zebrafish), border cells (Drosophila egg chamber), sprouting vessels, many epithelial and other cancer typesAny 2D and 3D ECM environment, resulting in cohesive sheets or 3D strands, tubes, clusters or amorphous massesIntact and stable cell–cell adhesions; coordination of multicellular leading edge protrusion and rear retraction; cell–cell communication during migrationCollective-to-single cell transitions (epithelial/collective-to-mesenchymal; collective-to-amoeboid)Hegerfeldt et al., 2002; Thiery, 2002; Alexander et al., 2008; Friedl and Gilmour, 2009
    Keratocyte-likeKeratinocytesObligate 2D surface or tissuePersistent gliding-type migration of spread-out cells with broad continuous leading lamella cadherin-based cell–cell junctionsNot knownKeren et al., 2008
Open in a separate windowAmoeboid migration commonly refers to the movement of rounded or ellipsoid cells that lack mature focal adhesions and stress fibers (Friedl et al., 2001; Lämmermann and Sixt, 2009). There are two subtypes of amoeboid movement. The first is the rounded, blebby migration of cells that do not adhere or pull on substrate but rather use a propulsive, pushing migration mode (Fackler and Grosse, 2008; Sanz-Moreno and Marshall, 2009). The second occurs in slightly more elongated amoeboid cells that generate actin-rich filopodia at the leading edge that engage in poorly defined, weak adhesive interaction with the substrate (Fig. 1; Yoshida and Soldati, 2006; Smith et al., 2007). In a special case of amoeboid movement, terminally matured nonadhesive dendritic cells produce dynamic actin-rich dendrites, instead of blebs, at their leading edge that cause these cell to become entangled with the ECM substrate during migration (Gunzer et al., 1997; Lämmermann et al., 2008). Individual cells with high levels of attachment and cytoskeletal contractility develop mesenchymal migration, which involves focalized cell–matrix interactions and movement in a fibroblast-like manner (Kaye et al., 1971; Maaser et al., 1999; Grinnell, 2008). The migration of individual cells that transiently form and resolve cell–cell contacts while moving along a common track is termed chain migration or cell streaming (Davis and Trinkaus, 1981; Teddy and Kulesa, 2004). Finally, the maintenance of stringent cell–cell adhesions can lead to partial or complete silencing of migration activity in cells inside a group yet supports cytoskeletal activity at outward edges or at basal cell–substrate contacts. The resulting collective migration occurs in the form of multicellular tubes, strands, irregularly shaped masses, or sheets (Vaughan and Trinkaus, 1966; Friedl et al., 1995; Farooqui and Fenteany, 2005).Most migration modes, although they can be observed in (mostly experimental) 2D environments, occur in vivo in the context of 3D tissue environments (Even-Ram and Yamada, 2005). Conversely, in vivo, some migration modes are dedicated exclusively to 2D environments. Epithelial keratocytes and keratinocytes migrate across flat 2D substrate using rapid spread-out cell gliding (Keren et al., 2008) that, if cell–cell junctions between the cells remain intact, form a collectively migrating 2D cell sheet (Vaughan and Trinkaus, 1966; Farooqui and Fenteany, 2005). In different cell types, these modes of migration are associated with different efficiencies yielding varying migration speeds, such as the fast migratory scanning of single leukocytes, the relatively slow invasive migration of fibroblasts into provisorial wound matrix, or, at the slowest end, the collective migration during organ formation (Friedl et al., 1998b).Single-cell and collective migration modes serve mutually exclusive purposes during morphogenesis, tissue regeneration, and in pathological conditions. Collective cell migration is essential in building, shaping, and remodeling complex tissues and tissue compartments, such as epithelia, ducts, glands, and vessels, but also contributes to cancer progression by local invasion (Alexander et al., 2008; Friedl and Gilmour, 2009). In contrast, single-cell migration allows cells either to cover local distances and integrate into tissues, such as neural crest cell migration, or to move from one location in the body to another and fulfill effector functions, such as immune cell trafficking (Friedl, 2004; Teddy and Kulesa, 2004; Lämmermann and Sixt, 2009). The latter process is recapitulated during cancer metastasis to distant sites (Thiery, 2002). Although not all molecular determinants of each migration mode are fully understood, some key parameters have been identified as “checkpoints” to either maintain a given migration type, or, by an increase or decrease of activity, initiate transitions.

Determinants of cell migration

The common process underlying all migration modes of nucleated mammalian cells is polarized actomyosin-driven shape change of the cell body (Lauffenburger and Horwitz, 1996; Ridley et al., 2003; Keren et al., 2008). This basic program is regulated and “shaped” by several distinct yet interdependent physical and molecular parameters of the tissue and the cell itself that together determine how a cell migrates (Fig. 2). The extracellular environment strongly impacts migration type and efficiency by providing ECM ligands of different macromolecular and structural organization, which includes dimension, density, stiffness, and orientation. In response to environmental determinants, the actomyosin cytoskeleton adapts in a dynamic manner and generates different geometries in space and time, ranging from flat and spread out to roundish, elongated, or multipolar shapes (Grinnell, 2008; Keren et al., 2008). To transmit actomyosin-driven forces to surrounding tissue structures, the cell either develops actin-polymerization–driven protrusions that bind to adhesion sites of the tissue through adhesion receptors (Yamada et al., 2003), or it utilizes poorly adhesive intercalation and propulsion (Paluch et al., 2006a). In both cases, subsequent to leading edge protrusion, actomyosin contraction leads to retraction of the cell rear and translocation of the cell body (Paluch et al., 2006a; Lämmermann and Sixt, 2009). The cyclic repetition of protrusion, interaction with the extracellular environment, and retraction of the cell rear result in cell movement that, depending on the molecular repertoire of the cell, yields distinct migratory modes (Lauffenburger and Horwitz, 1996; Friedl and Wolf, 2009). Additional parameters impacting the type and efficiency of cell migration are the availability of surface proteases that remodel the surrounding tissue (Wolf and Friedl, 2009), and whether the cells retain stringent, loose, or no cell–cell junctions (Friedl and Gilmour, 2009).Open in a separate windowFigure 2.The tuning model of cell migration. An integrated multiscale model to combine multiple interdependent parameters that impact migration mode. Each parameter is experimentally testable individually; however, in most cases they are interconnected with others (see text for details). Approximated parameter profiles of selected migration modes are indicated (colored lines). Modulation by increasing or decreasing the magnitude of any parameter may impact the resulting migration mode as well as the input strength of coregulated parameters. The format of the tuning model mimics the popular display of a graphic equalizer, which is integral to modern media display programs (e.g., Windows Media Player or QuickTime); the graphic interface serves to adjust the intensity of different wavelengths of the phono output independently to modify the sound profile.

ECM determinants

The ECM provides a structural and molecular frame for the moving cell body and thereby impacts the mode and efficiency of cell migration.

ECM dimension.

Extracellular tissue structures encountered by migrating cells are either flat 2D sheets or 3D tissue networks. Cell migration across 2D surfaces occurs during reepithelialization of wounds or the scanning of leukocytes along the inner blood vessel wall or inner epithelial surfaces (Farooqui and Fenteany, 2005). Hallmarks of 2D migration are the requirement of unilateral adhesion to the substrate, which provides stable-enough but transient attachment; a flattened, spread-out cell morphology guided by a leading lamellipod; and, due to the flat geometry of the substrate, a largely barrier-free migration (Ridley et al., 2003; Farooqui and Fenteany, 2005; Keren et al., 2008; Vitorino and Meyer, 2008). In contrast, when cells move through 3D interstitial tissue consisting of a network of interwoven collagen fibers, which impose space limitations against the moving cell body, their morphology undergoes characteristic changes. First, spread-out morphology is abandoned in favor of a spindle-like shape; second, instead of lamellipodia formation, with its unilateral polarization to the underlying substrate, leading edge protrusion occurs by formation of thin tiplike cylindrical pseudopodia that orient in three dimensions; and third, the cell either deforms its shape to accommodate small tissue gaps or executes remodeling of the ECM structure by pericellular proteolysis (Maaser et al., 1999; Wolf et al., 2003a; Jiang and Grinnell, 2005).

ECM density and gap size.

In vivo, interstitial tissues greatly vary in structural organization, such as collagen content, fibrillar texture, fiber bundle thickness, and interfiber porosity. In vivo, migration efficiency is optimal at pore diameters that match or range slightly below the diameter of polarized cells. If the tissue gaps exceed the cell size, migration rates decrease (Haston et al., 1982; Harley et al., 2008) because of a loss of most cell–fiber interactions until only very few or even a single fiber remain engaged with the cell body; the latter is termed “1D” migration (Doyle et al., 2009). Conversely, if pores range below the cell diameter, cells slow down and eventually may become trapped due to the physical hindrance (unpublished data; Haston et al., 1982; Harley et al., 2008). In response to extracellular confinement, migrating cells elongate to a spindle-like shape and thereby stretch and reduce their cell diameter, whereas large pore sizes favor cell rounding, a hallmark of amoeboid migration (unpublished data; Fig. 2).The deformability of the cell and its most rigid compartment, the nucleus, is controlled by nuclear lamins A/C, which mechanically stabilize the nuclear membrane and potentially impact the minimum tissue gaps that can be transmigrated (Lammerding et al., 2006; Dahl et al., 2008). Besides shape adaptation, cells that can proteolytically cleave ECM structures counteract physical hindrance by enlarging pores and forming trails of variable caliber so they match their own diameters (Wolf et al., 2007). Thus, the ability of the cell to deform relative to the available space and to remodel tissues through proteolysis determines both the mode and efficiency of migration in 3D ECM.

Stiffness.

ECM stiffness (synonymous with rigidity) or elasticity (synonymous with pliability), which can be measured as elastic modulus, depends on molecular properties of the tissue, including collagen content, fiber thickness, and the extent of intrafibrillar cross-links, which define the stability and deformability of the tissue scaffold (Shoulders and Raines, 2009). Cells detect matrix rigidity via integrin-mediated adhesions and downstream mechanosensor protein signaling (i.e., via talin and p130CAS; Giannone and Sheetz, 2006). Increased substrate stiffness reinforces cell protrusions at outward edges so that focal adhesions form and become reinforced by RhoA-mediated actomyosin contraction, ultimately leading to cell spreading, the generation of high-traction force, and elongated cell movement (Peyton et al., 2008; Ulrich et al., 2009). Conversely, soft matrix does not reinforce focal adhesion formation and cytoskeletal contractility; rather, it supports cell rounding (Ulrich et al., 2009). Consequently, matrix rigidity stimulates directed cell migration, similar to chemotaxis, so that cells tend to migrate toward substrate of greater stiffness, a process termed durotaxis (Lo et al., 2000; Li et al., 2005; Isenberg et al., 2009).

Orientation.

Connective tissue comprises a range of physical textures, ranging from loose and random to highly aligned structures (Petrie et al., 2009; Wolf et al., 2009). All mobile cells show a tendency to align in parallel along oriented structural discontinuities, such as at interfaces of muscle fibers, blood vessels, or ECM fiber strands and patterns created by the cells themselves (Provenzano et al., 2008; Petrie et al., 2009). Contact guidance along such structures is mediated by mechanosensory integrins that, together with Rho/ROCK-mediated cytoskeletal stiffening, provide directional persistence (Dickinson et al., 1994; Provenzano et al., 2008; Petrie et al., 2009). Although aligned fiber orientation in collagen-rich ECM does not seemingly impact cell shape (Provenzano et al., 2008), it favors multicellular streaming in chainlike patterns in 3D tissue (Friedl and Wolf, 2009) and migration of 2D cell sheets along tissue clefts (unpublished data).In summary, different ECM environments provide an array of interconnected input parameters that modulate cell adhesion and cytoskeletal organization, and directly impact cell shape, guidance, and mode of migration.

Cell determinants

Cell–cell adhesion.

A key determinant of how cells move is whether cell–cell junctions are retained or not, resulting in either collective or single-cell migration, respectively (Vitorino and Meyer, 2008; Friedl and Gilmour, 2009). Cell–cell adhesion is mainly mediated by cadherins, including E-cadherin in epithelial cells, VE-cadherin in endothelial cells, and N-cadherin in stromal cells (Ewald et al., 2008; Vitorino and Meyer, 2008; Friedl and Gilmour, 2009). As opposed to individually migrating cells, during collective migration, the rear of the front cell retains intact cell–cell junctions to the successor cell, thereby mechanically holding the cells together and augmenting the efficiency of paracrine cell–cell signaling and multicellular coordination (Fig. 1). Coordinated cycles of protrusion and rear retraction of the front cells as well as of cells inside the group that engage with underlying substrate lead to movement as a multicellular unit (Farooqui and Fenteany, 2005; Blanchard et al., 2009). If cell–cell junctions are intermittent or less stable, multicellular streaming in a loose tail-to-head fashion results in the coordinated but individual migration of many cells through the tissue, with repetitive short-lived contacts between cells that are resolved and reestablished upon further migration (Fig. 1; Teddy and Kulesa, 2004). Lastly, if cell–cell contacts are absent, cells move independently in both speed and direction (Hegerfeldt et al., 2002). Thus, the presence of stable or transient cell–cell junctions, or their absence, determines whether collective translocation, cell streaming, or single-cell migration, respectively, is being generated.

Cell–matrix adhesion.

Cell adhesions to ECM ligands are predominantly generated by integrins via coupling to cytoskeletal and signaling proteins. The strength and turnover rates of cell attachments to the extracellular environment determine which cell shapes and forces are being generated during migration (Ridley et al., 2003). Distinct cell types use adhesive strength over different magnitudes, ranging from strong adhesion by stromal fibroblasts or myoblasts (Huttenlocher et al., 1996), to moderate adhesion of epithelial and endothelial cells (Zhang et al., 2006; Schober et al., 2007), to weak adhesion forces of rapidly gliding fish keratocytes and crawling leukocytes (Friedl et al., 1998b; Keren et al., 2008; Lämmermann et al., 2008). High integrin expression levels are mandatory for high-attachment forces, but are also associated with relatively slow turnover of adhesion sites (Friedl et al., 1998b; Mc Henry et al., 2008) and, consequently, associated with slow migration (Palecek et al., 1997). As an underlying mechanism, integrins and downstream mechanotransducing adaptors, such as p130CAS, become activated with increased mechanical tension and, in turn, further strengthen focal adhesions and actin stress fiber formation (Tamada et al., 2004; Sawada et al., 2006). Strong cell–substrate adhesions thus promote cell contractility and the formation of elongated spread-out (2D) or spindle-shaped (3D) morphologies in many cell types, including fibroblasts, smooth muscle cells, and neoplastic cells (Lauffenburger and Horwitz, 1996; Friedl et al., 1998b; Maaser et al., 1999; Jiang and Grinnell, 2005).If cell adhesion is reduced to a moderate or low level, such as by interfering with the integrin-talin axis, focal adhesions and stress fibers do not form or do not reach full maturation (Zhang et al., 2008). As a consequence, the cells convert to a less elongated or spread-out morphology, generate smaller lamellipodia and pseudopodia, and transmit limited adhesion strength toward the substrate (Zhang et al., 2008). Rapidly moving lymphocytes and neutrophils that still adhere to ECM and other ligands but do not form focal adhesions or stress fibers constitutively use the pseudopodal amoeboid type of movement (Friedl et al., 1998a; Smith et al., 2007).At the very low end of cell adhesion strength, cells are unable to form unilateral attachments to 2D ECM substrate and thus fail to spread out, form lamellipodia, and move, whereas in a 3D environment, they move by amoeboid blebbing or dendritic intercalation (Haston et al., 1982; Fackler and Grosse, 2008; Lämmermann and Sixt, 2009). Given such low adhesion capability, the mechanisms that generate force in this blebby (or dendritic) amoeboid translocation remain to be shown. Likely, the irregular cell shape maintained by cortical actin provides high cytoskeletal rigidity locally, which allows mechanical intercalation between anterior parts of the cell with the surrounding tissue while the rear part of the cell retracts (Blaser et al., 2006; Paluch et al., 2006a; Lämmermann and Sixt, 2009).

Cell protrusion and rounding.

Cell protrusions control leading edge dynamics and the migration mode in at least two distinct ways. First, the protrusion of pseudopodia, filopodia, and lamellipodia that adhere to cell and ECM substrates is directed by the small GTPases Rac and Cdc42 (Nobes and Hall, 1999; Sanz-Moreno and Marshall, 2009). Consequently, high Rac activity conveys leading edge extension, elongated morphology, focal integrin engagement, and mesenchymal migration (Nobes and Hall, 1999; Sahai and Marshall, 2003; Sanz-Moreno et al., 2008). Second, bleb-like protrusions that contain cortical actin filaments are nonadhesive or poorly adhesive but contribute to lateral anchoring (“elbowing”) of the cell to tissue structures during actomyosin-mediated rear retraction (Paluch et al., 2006a,b; Fackler and Grosse, 2008). In most cells, Rac-mediated protrusion of the leading edge is counterbalanced by Rho/ROCK signaling, which controls actomyosin-mediated retraction of the trailing edge. Together, they form a cyclic balance in distinct regions of the cell and contribute, concurrently, to the migration cycle (Ridley et al., 2003; Sanz-Moreno and Marshall, 2009). High Rac activity generates cell elongation and mesenchymal migration, whereas active Rho in the presence of little or no Rac activity supports rounded cell shapes associated with amoeboid pseudopodal or blebbing migration, respectively (Sahai and Marshall, 2003; Sanz-Moreno et al., 2008). Besides inducing cell protrusions, active Rac negatively regulates Rho/ROCK signaling and inhibits cell rounding, whereas active Rho/ROCK limits Rac, which inhibits cell extension and elongation (Sanz-Moreno et al., 2008).The formation and elongation of cell protrusions during migration are further controlled by tubulins. Posttranslational tubulin acetylation supports high microtubule stability and is associated with mesenchymal movement, whereas microtubules composed of deacetylated tubulin are subject to enhanced depolymerization by the microtubule-destabilizing factor stathmin and therefore support a rounded, amoeboid migration mode (Piperno et al., 1987; Belletti et al., 2008; Berton et al., 2009). Whether tubulin stability dictates cell shape by modulating to the balance between Rac and Rho activity or by other mechanisms, such as delivery of cargo or a direct mechanical function, is unknown.

Mode of force generation.

The force required to move a cell body forward is generated by two principal and often interdependent physical mechanisms: cell propulsion, which leads to forward pushing of the cell body; or traction force generated by pulling of an ECM substrate. A phase of actin polymerization–driven forward pushing of the plasma membrane is indispensible for leading edge protrusion, so it is included in most migration types (Lauffenburger and Horwitz, 1996). In adhesive cells, pushing then leads to local adhesion, cytoskeletal anchorage, and, in a second phase, focal adhesion maturation and pulling on ECM substrate by actomyosin contraction (Ridley et al., 2003; Zhang et al., 2008). Pulling is proportional to adhesion strength and cytoskeletal contractility, such as in fibroblasts and myoblasts, to generate forces sufficient for substrate contraction (Beningo et al., 2001; Miron-Mendoza et al., 2008). In contrast, if leading edge protrusion is coupled to low adhesion force, amoeboid pseudopodal migration occurs at very low traction force, as in moving neutrophils (Smith et al., 2007; Wang, Y.-L., personal communication). On the very low end of adhesion and force generation, amoeboid blebbing cells tend to lack any attachment to 2D surfaces but rather float and oscillate on the spot (unpublished data; Paluch et al., 2006a). However, if included in a loose 3D ECM, such as a collagen matrix or matrigel, blebby cells that are deficient in pseudopodia or filopodia are still able to connect to the 3D substrate and generate movement, despite negligible attachment forces (Blaser et al., 2006; Sanz-Moreno et al., 2008). Thus, whereas mesenchymal migration depends on alternating pushing/pulling cycles, amoeboid migration is mechanically equally complex and comprises stronger pushing combined with a small or completely absent phase of adhesive pulling of the substrate.

Protease functions.

Depending on the deformability of the migrating cell and the size of gaps and trails available in the 3D tissue, cells proteolytically remodel surrounding ECM and generate gaps, a hallmark of mesenchymal migration; otherwise, they move without engaging proteases by filling available spaces with their cell body (Friedl and Wolf, 2003a, 2009). In interstitial tissues, MT1-MMP is rate-limiting for collagen degradation, as it executes pericellular proteolysis of collagen fibers that physically impede the moving cell (Wolf et al., 2007; Sabeh et al., 2009). After cleavage, collagen fibers become displaced and realigned, which generates tubelike matrix gaps and trails of least resistance (Friedl and Wolf, 2008). In collagen-rich interstitial tissue, MT1-MMP is further involved in the remodeling of already existing trails to even larger macrotracks, which then accommodate the collective invasion of multicellular strands (Wolf et al., 2007).In contrast to mesenchymal cells that are usually large, smaller amoeboid leukocytes employ much faster movement that lacks signs of pericellular proteolysis of the 3D interstitial substrate (Friedl and Wolf, 2003a). A mechanism of coping with narrow trails is cell deformation and squeezing through the pores so that extracellular structures imprint into the cell body and form local zones of cell compression (Wolf et al., 2003b). If tissue densities are high, such as in basement membranes or dense connective tissue, inhibition of pericellular proteolysis cannot be compensated by shape change; instead, cell bodies get stuck in narrow pores (Sabeh et al., 2004, 2009). Likewise, if proteolytic macropatterning is prevented by protease inhibition, collective cell invasion is ablated and only individual amoeboid dissemination persists (Wolf et al., 2007). Thus, proteolytic ECM remodeling is obligatory in tissues in which cell caliber and deformability fail to match available gaps and trails.

The tuning model

Because of its physical and molecular modularity, cell migration must be viewed as a consequence of a continuum of states that are determined by cell mechanics and signaling events. These cellular properties are integrated by the cell or cell groups in a given tissue environment. The tuning model predicts that several parameters simultaneously control how a cell migrates and that their combined magnitudes impact which migration type a cell adopts (Fig. 2). With the exception of ECM dimension, which is either 2D or 3D, all other parameters are scalar; i.e., they can be absent or at low, intermediate, or high levels. Therefore, these parameters are assumed to be tunable and thereby control the migration mode and efficiency in a continuous rather than a discrete “on” or “off” manner. By increasing or decreasing their input, they “tune” how moving cells polarize and engage with encountered tissue substrate. Because all parameters act concurrently but at a different strength, each parameter profile (Fig. 2, colored lines) then generates a different type of migration. Whereas most molecular studies tend to address isolated parameters, the tuning model integrates several denominators in context and may help to understand cell migration as a multimodal cell function.Each component, although experimentally amenable as an individual parameter, is interdependent and positively or negatively coregulated with other determinants. The density of fibrillar ECM is positively interconnected with stiffness and inversely proportional to pore size, so alterations of either parameter impacts the overall tissue geometry (unpublished data). Accordingly, integrin-mediated cell attachment to a deformable yet rigid substrate, but not to a soft substrate, enhances substrate tension and stiffness, which reinforces Rho-mediated traction force generation (Paszek et al., 2005; Peyton et al., 2008; Ulrich et al., 2009). Likewise, traction force generation requires sufficient adhesion mediated by integrins, some Rac-mediated protrusion, and Rho-mediated cytoskeletal contraction (Rhee and Grinnell, 2006). The physical tissue geometry is interdependent with protease acitivity of the cells; consequently, collective migration in 3D tissue depends on sufficiently high a priori porosity or the cell-mediated proteolytic generation of macrotracks (Wolf et al., 2007). Therefore, alteration of a given parameter has likely consequences for other interconnected determinants.

Plasticity: tuning the mode of migration

At a given differentiation state, each cell type preferentially employs a particular “default” migration type, such as leukocytes using amoeboid migration, stromal cells moving by a mesenchymal mode, or epithelial cell sheets moving collectively (Friedl, 2004). However, in recent years, it has become clear that naturally occurring or experimentally induced modifications of either the environment or cell properties may result in striking adaptation reactions that alter the migration mode rather than abrogating migration per se. Because any parameter may become altered in the course of migration—such as the transition from dense to loose connective tissue, modulation of adhesion receptor expression, or the availability of cytoskeletal adaptor proteins due to altered gene expression—each alteration of parameter may prompt such secondary alteration of migration mode.Because cell–cell junctions can form de novo and resolve again, individual and collective migration modes are interconvertible (Friedl and Gilmour, 2009). If multicellular cohesion is weakened by the down-modulation of cell–cell junctions, individual cells detach from the multicellular unit which, dependent on the molecular repertoire and environment encountered, disseminate individually. Epithelial-to-mesenchymal transition is involved in many developmental processes and in invasive cancers, and leads to the delamination of spindle-shaped cells that use integrin-mediated force generation for tissue invasion either as single cells or by multicellular streaming (Thiery, 2002; Carmona-Fontaine et al., 2008). Collective-to-amoeboid transition occurs when the detached individual cells disseminate using amoeboid migration (Hegerfeldt et al., 2002; Wolf et al., 2007). Conversely, if individually moving cells up-regulate cell–cell adhesion molecules, then cell aggregation leads to individual-to-collective transition (Thiery, 2002).A central pathway controlling the interconversion between mesenchymal and rounded, amoeboid migration is the balance between Rac and Rho signaling (Sahai and Marshall, 2003). In many experimental examples, mesenchymal-to-amoeboid transitions depend, directly or indirectly, on pathways that weaken Rac and/or strengthen Rho/ROCK signaling (Sanz-Moreno et al., 2008; Paňková et al., 2009; Sanz-Moreno and Marshall, 2009). Thus, upstream pathways that suppress Rac activity induce this conversion, including the activation of the GTPase-activating protein (GAP) ARHGAP22, which directly inhibits Rac (Sanz-Moreno et al., 2008); inhibition of the Rac-activating guanine nucleotide exchange factor DOCK3/NEDD9 or of the down-stream Rac effector and Rho inhibitor WAVE-II (Sanz-Moreno et al., 2008); interference with Rab5-mediated endocytosis and recycling of Rac to cell protrusions (Palamidessi et al., 2008); and the inhibition of E3 ubiquitin ligase Smurf1, which enzymatically degrades Rho near the leading edge and thereby secures the dominance of Rac in cell protrusions (Sahai et al., 2007). Likewise, inhibition of chemokine-meditated Rac activation favors amoeboid movement in otherwise mesenchymal cells (Gérard et al., 2007). Pathways that activate Rho lead to mesenchymal-to-amoeboid transition, including inhibition of negative Rho regulators (e.g., p90RhoGAP) through an indirect, reactive oxygen species-dependent mechanism (Nimnual et al., 2003) or the activation of Ephrin2A receptor tyrosine kinase signaling, which indirectly activates Rho (Parri et al., 2009).Alteration of adhesion force by modulating integrin function leads to similar plasticity. The transition from amoeboid myeloid precursor cells to adhesive elongated and contractile macrophages is initiated by the up-regulation and activation of β1, β2, and β3 integrins (McNally and Anderson, 2002). Conversely, mesenchymal cells convert to amoeboid movement in experimental 3D matrices after limiting pathways that control focal adhesion formation, including inhibition of β1 integrin–mediated adhesion or of the tyrosine kinase cSrc (Carragher et al., 2006; Zaman et al., 2006). Plasticity imposed by altered tissue architecture occurs when cells transit from a 2D interface into 3D tissue. The initial spread-out or cuboidal 2D phenotype then converts toward a spindle-shaped mesenchymal phenotype with vertical penetration and migration into the 3D matrix (Alt-Holland et al., 2008). Lastly, in loose interstitial tissues with gaps that accommodate the cell body, the inhibition of surface protease activity causes a transition from protease-dependent mesenchymal migration to amoeboid migration involving shape change and squeezing without tissue remodeling (Wolf et al., 2003a, 2007).Thus, alterations of cell–cell and cell–matrix adhesion, cytoskeletal signaling and mechanics, and protease function determine whether and how cells switch between distinct migration modes. These transitions of migration mode are best studied for cancer cells and likely contribute to the metastatic cascade (Friedl and Wolf, 2003b; Sanz-Moreno and Marshall, 2009), yet they are also relevant to cell migration and function in physiological contexts, such as the delamination of cells during morphogenesis and the distribution of stem cells or leukocytes in tissues and organs (Blaser et al., 2006; Lämmermann and Sixt, 2009).

Outlook

The multiparameter tuning model integrates observations from many different cell types and experimental models. The model thus may be helpful to understand and experimentally test the adaptability of cell movement and its consequence for tissue formation and remodeling, particularly in morphogenesis and cancer metastasis. The model may further be a useful starting point for computational modeling of cell migration in different contexts. Although the parameters and migration modes discussed here are best established for interstitial migration of cells in fibrillar collagen-rich tissues, they likely fail to sufficiently represent the movement of other cell types and tissue contexts. This may be the case particularly for cells of neural origin that predominantly move along scaffold tracks formed by other cells, rather than ECM, or cell trafficking across basement membrane during transendothelial migration or the early invasion of epithelial cancer. Likewise, complex movements in ductal gland or vessel formation represent special cases with complex topography, such as lumen formation and deposition of a basement membrane, which may require the inclusion of additional modules. Besides integrin-mediated adhesion structures, special cases of cell–substrate interaction include cadherin- or ephrin-based cell–cell junctions that guide cell migration along cell scaffolds, and podosomes and invadopodia that degrade ECM underneath the cell body but not at leading edges. The contribution of these structures to force generation and the mode of migration remain to be established and, potentially, included in the model. Ultimately, although each parameter has its own contribution to how efficiently cells migrate, the model still lacks prioritization; that is, the importance of each input parameter relative to others still remains undefined. Therefore, future wet-laboratory and computational studies will not only have to integrate additional or exclude existing determinants for special migration modes and contexts, but they also should take coregulated synergistic or antagonistic multiparameter modules into account.  相似文献   

20.
The cell biology of disease: The cellular and molecular basis for malaria parasite invasion of the human red blood cell     
Alan F. Cowman  Drew Berry  Jake Baum 《The Journal of cell biology》2012,198(6):961-971
  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