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 78 毫秒
1.
2.
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   

3.
Even with the assistance of many cellular factors, a significant fraction of newly synthesized proteins ends up misfolded. Cells evolved protein quality control systems to ensure that these potentially toxic species are detected and eliminated. The best characterized of these pathways, the ER-associated protein degradation (ERAD), monitors the folding of membrane and secretory proteins whose biogenesis takes place in the endoplasmic reticulum (ER). There is also increasing evidence that ERAD controls other ER-related functions through regulated degradation of certain folded ER proteins, further highlighting the role of ERAD in cellular homeostasis.Newly synthesized membrane and secreted proteins enter the ER in an unfolded state through a protein-conducting channel named the translocon (Rapoport, 2007). In the ER, a myriad of chaperones and modifying enzymes assist their membrane integration and folding. In many cases folding involves post-translational modifications, such as glycosylation or disulfide bond formation (Braakman and Hebert, 2013). At this stage many proteins are also assembled into multisubunit complexes with defined stoichiometries. As newly synthesized proteins reach a native conformation, they leave the ER to perform their function elsewhere; either along the secretory pathway or outside of the cell.Despite all the resources dedicated to protein folding, a significant fraction of newly synthesized polypeptides entering the ER fails to acquire a native conformation (Hartl and Hayer-Hartl, 2009). The degree of misfolding of these proteins varies considerably and can have several causes such as mutations, substoichiometric amounts of a binding partner, or merely a shortage of chaperone availability. In most cases, the misfolded molecules are retained in the ER and eventually become substrates of the ER-associated protein degradation (ERAD), a collection of quality-control mechanisms that clears the ER from these potentially harmful species. Inactivation of ERAD results in the accumulation of misfolded proteins in the lumen and membrane of the ER, a condition known as ER stress that is common to several diseases (Walter and Ron, 2011). For this reason, ERAD plays a key role in ER homeostasis across eukaryotes. Genetic ablation of a number of ERAD components leads to embryonic lethality in mice, also highlighting the importance of this process in cellular and organismal homeostasis (Yagishita et al., 2005; Francisco et al., 2010; Eura et al., 2012). Whether this essential function of ERAD during mouse development is due to its role in the degradation of misfolded proteins remains to be determined.Certain folded, perfectly active proteins are also targeted by ERAD. However, their degradation is highly regulated and only occurs in the presence of a specific signal. The best-characterized regulated substrate is the 3-hydroxy-3-methylglutaryl acetyl-coenzyme-A reductase (HMGR), a key enzyme in sterol biosynthesis (Gil et al., 1985; Hampton et al., 1996; Bays et al., 2001a; Song et al., 2005). Both in yeast and in mammals, HMGR degradation by ERAD is part of a feedback inhibition system critical for sterol homeostasis. Interestingly, another enzyme of the sterol biosynthetic pathway, squalene monooxygenase (Erg1 in yeast and SQLE in mammals), was recently identified as a regulated ERAD substrate (Foresti et al., 2013). The degradation of Erg1/SQLE by ERAD is again part of a feedback inhibition system to prevent the accumulation of intermediate sterol metabolites, which are toxic for cells (Foresti et al., 2013). Recent evidence shows that regulation of the synthesis of sterols and other sterol-derived metabolites by ERAD is also present in plants (Doblas et al., 2013; Pollier et al., 2013). This evolutionarily conserved role of ERAD in sterol regulation might have been one of its primordial functions.The ERAD machinery is also exploited by certain viruses to degrade host proteins thereby escaping immune surveillance. Well characterized examples are the degradation of newly synthesized major histocompatibility complex class I (MHC I) heavy chain (Wiertz et al., 1996a) or CD4 molecules by the human cytomegalovirus or the immunodeficiency virus (HIV; Fujita et al., 1997; Schubert et al., 1998), respectively. Moreover, some bacterial toxins, such as cholera, and viruses, like simian virus 40 (SV40), travel to the ER retrogradely through the secretory pathway. At the ER these toxins and viruses exploit ERAD components to reach the cytosol, where ultimately they will act (Tsai et al., 2001; Schelhaas et al., 2007; Bernardi et al., 2008).Finally, ERAD components are also involved in the turnover of several soluble proteins in the cytoplasm and the nucleus of cells (Swanson et al., 2001; Ravid et al., 2006; Yamasaki et al., 2007). Most of these cases, however, involve only a subset of the ERAD steps and components. In sum, although a complete repertoire of substrates is not available, it is clear that misfolded proteins are not the exclusive clients of ERAD.

ERAD, linking ER quality control to cytoplasmic protein degradation

The earliest evidence for protein quality control at the ER came from observations that unassembled subunits of the T cell receptor were rapidly degraded in the cells (Lippincott-Schwartz et al., 1988). This degradation occurred independently of lysosomal proteases, leading to the proposal that the ER itself would house some uncharacterized proteolytic activity toward misfolded proteins. Then a landmark study in yeast showed that the degradation of a short-lived misfolded ER membrane protein was blocked in cells lacking Ubc6, a component of the ubiquitin conjugation machinery (Sommer and Jentsch, 1993). The ubiquitin system mediates the covalent attachment of ubiquitin, a small 76–amino acid protein, to target proteins in the cytoplasm by the sequential action of activating (E1), conjugating (E2), and ligase (E3) enzymes (Pickart, 2001). Ubiquitin-modified proteins are then recognized and degraded by the proteasome. The involvement of the ubiquitin–proteasome system in ER protein quality control was confirmed by studies on the degradation of mutant and wild-type cystic fibrosis transmembrane conductance regulator (CFTR), a large membrane protein with a complicated folding process (Jensen et al., 1995; Ward et al., 1995). Inhibition of proteasome function led to accumulation of CFTR molecules, and interestingly, a significant fraction of these was detected as ubiquitin conjugates (Jensen et al., 1995; Ward et al., 1995). Soon after, it became clear that a similar mechanism could also account for the degradation of luminal misfolded proteins such as CPY*, a mutant version of the yeast vacuolar carboxypeptidase Y and a prototype ERAD substrate (Hiller et al., 1996). Together, these papers demonstrated that aberrant proteins in the lumen and membrane of the ER are degraded in the cytoplasm where the components of the ubiquitin–proteasome system reside.

Ubiquitin ligase complexes: The hubs in ERAD

Subsequent genetic and biochemical studies, primarily in budding yeast but also in mammalian cells, identified many ERAD components and led to a general understanding of the organization of the pathway. An important realization was that the “one-size-fits-all” model does not apply to ERAD and that this pathway encompasses multiple branches with distinct specificity for different classes of misfolded proteins (Taxis et al., 2003; Vashist and Ng, 2004; Carvalho et al., 2006; Bernasconi et al., 2010; Christianson et al., 2012). However, irrespective of the branch, the same sequence of events leads to the degradation of all ERAD substrates (Fig. 1 A). The first step is the recognition of a substrate in the crowded ER environment. Then the substrate is transported across the ER lipid bilayer back into the cytoplasm, a step known as retrotranslocation. On the cytosolic side of the ER membrane, the substrate is ubiquitinated by a membrane-associated ubiquitin ligase (or E3 ligase). Subsequently, the ubiquitinated substrate is extracted from the membrane in an ATP-dependent manner and released in the cytoplasm for degradation by the proteasome. The execution of these steps is coordinated by a membrane-embedded protein complex named after the E3 ligase at its core. The canonical E3 ligases involved in ERAD are themselves multispanning membrane proteins, in which the RING (really interesting new gene) domain responsible for the ligase activity is in the cytoplasm. These E3 ligase complexes are best characterized in yeast (Fig. 1 B and Swanson et al., 2001) and Hrd1 (Bordallo et al., 1998; Bays et al., 2001a) assemble into the Doa10 and the Hrd1 complexes, respectively, each responsible for the degradation of a class of ERAD substrates (Carvalho et al., 2006).Open in a separate windowFigure 1.The different steps and branches in ERAD. (A) The events defining the ERAD of a generic luminal substrate with a misfolded domain (red star). Substrate recognition, retrotranslocation, and ubiquitination are coordinated by a membrane-embedded E3 ligase complex. Ubiquitin is depicted as small circles. (B) The E3 ligase complexes involved in ERAD in S. cerevisiae and their substrate specificities. ER proteins with a misfolded domain in the cytoplasm (ERAD-C substrates) are degraded via the Doa10 complex. Proteins with luminal (ERAD-L) or intramembrane (ERAD-M) misfolded domains are degraded via the Hrd1 complex. Misfolded domains on proteins are indicated by a red star. The Cdc48 cofactors Npl4 and Ufd1 are depicted as N and U, respectively.

Table 1.

Components of the yeast E3 ligase complexes and their mammalian counterparts
ComponentFunctionMammalian homologue
Hrd1 complex
Hrd1E3 ligase activity/retrotranslocation?HRD1, gp78
Hrd3Substrate recognition, Hrd1 stabilitySEL1
Yos9Substrate recognitionOS9, XTP3-B
Kar2Chaperone activity, substrate recognitionBip
Usa1Hrd1 and Der1 oligomerizationHERP
Der1Recognition/transfer of substrate to Hrd1/retrotranslocation?Derlin-1, -2, -3
Doa10 complex
Doa10E3 ligase activityTEB4
Ubc6E2 ubiquitin-conjugating activityUbc6, Ubc6e
Common to Hrd1 and Doa10 complexes
Ubc7E2 ubiquitin-conjugating activityUBE2G1, UBE2G2
Cue1Recruitment and activation of Ubc7
Ubx2Membrane-recruiting factor for Cdc48UBXD8
Cdc48Substrate retrotranslocation and membrane extractionp97/VCP
Npl4Cdc48 cofactorNPL4
Ufd1Cdc48 cofactorUFD1
Open in a separate windowBased on the analysis of a few model substrates, the E3 ligase complex specificity appears to be determined by the location of the misfolded lesion on a substrate relative to the ER membrane: proteins with misfolded domains in the cytoplasmic side of the membrane (ERAD-C substrates) are degraded via the Doa10 complex; proteins with luminal (ERAD-L substrates) or intramembrane (ERAD-M substrates) misfolded domains are targeted to the Hrd1 complex (Fig. 1 B and Taxis et al., 2003; Vashist and Ng, 2004; Carvalho et al., 2006). Factors involved in substrate recognition are unique to the E3 ligase complex and likely determine the substrate specificity of each ERAD branch. On the other hand, the components that act at late steps of ERAD, such as the Cdc48 ATPase complex (p97 in mammals) required for membrane extraction of ubiquitinated substrates (Bays et al., 2001b; Ye et al., 2001; Jarosch et al., 2002; Rabinovich et al., 2002), are common to both E3 ligase complexes.In mammalian cells the best-studied E3 ligases are Hrd1 and Gp78 (Schulze et al., 2005; Mueller et al., 2008; Bernasconi et al., 2010; Christianson et al., 2012; Burr et al., 2013). Several more E3 ligases have been implicated in ERAD in mammalian cells (such as Rma1/Rnf5, Trc8, Rfp2, Rnf170, and Rnf185) but these are still poorly characterized. Only few substrates are known for each ligase and a preference for particular ERAD substrate classes has been difficult to infer (Claessen et al., 2012).

How are ERAD substrates recognized?

Recognition of misfolded proteins.

The commitment to degradation by ERAD occurs at the level of substrate recognition; therefore, this step needs to be tightly controlled. Inefficient detection of misfolded proteins leads to their accumulation, ultimately affecting cell function (Travers et al., 2000; Jonikas et al., 2009). On the other hand, overactive ERAD would likely have its cost, with the degradation of significant amounts of folding intermediates. For this reason, substrate recognition by ERAD has to be finely balanced. This is a complex task considering the ER environment, in which a complete spectrum of protein species coexist, from newly synthesized unfolded molecules to fully folded proteins.Folding intermediates and terminally misfolded proteins share structural similarities, for example the exposure of hydrophobic patches that are normally hidden once proteins acquire the native structure. These molecules are kept in a soluble state by binding to chaperones such as the ER Hsp70 (Kar2 in yeast and BiP in mammals), which are essential for the folding of newly synthesized polypeptides as well as for the disposal of misfolded proteins. However, chaperones on their own do not appear to determine the fate of their clients. Instead, recognition factors that are part of the E3 ligase complexes play a major role in ERAD substrate selection. For example, the recognition of ERAD-L substrates requires the luminal components of the Hrd1 complex Hrd3, Kar2 and, in the case of glycosylated substrates, the lectin Yos9 (Plemper et al., 1997, 1999; Bhamidipati et al., 2005; Kim et al., 2005; Szathmary et al., 2005; Carvalho et al., 2006; Denic et al., 2006; Gauss et al., 2006a). The yeast derlin Der1, a membrane protein of the Hrd1 complex, might also be involved in ERAD-L substrate recognition (Gauss et al., 2006b; Stanley et al., 2011). Moreover, certain ERAD-M substrates appear to be recognized directly by the E3 ligase Hrd1 (Sato et al., 2009). However, the features recognized on the misfolded proteins by these ERAD factors are largely unknown.An informative exception is the recognition of luminal misfolded N-linked glycoproteins in Saccharomyces cerevisiae (Fig. 2 A). As they enter the ER lumen, proteins are often modified at asparagine residues (in the context of the N-X-S/T sequence) with a well-defined, branched glycan moiety made up of three glucose, nine mannose, and two N-acetylglucosamine residues, Glc3–Man9–GlcNAc2 (Fig. 2 A; Braakman and Hebert, 2013). This N-linked glycan is subsequently trimmed by several enzymes. Early glycan-processing enzymes such as glucosidases lead to the binding of lectins that facilitate the folding of the newly synthesized proteins. In contrast, late acting enzymes, such as the mannosidase Htm1, trigger the binding of a different lectin that engages the protein in ERAD (Jakob et al., 2001; Quan et al., 2008; Clerc et al., 2009). This difference in the kinetics of the glycan-trimming enzymes provides an opportunity for newly synthesized proteins to acquire the native conformation and traffic beyond the ER (Fig. 2 A). A long ER residency, indicative of folding problems, results in the processing of the misfolded glycoproteins by Htm1, which generates a biochemical mark (α1,6-linked mannose) decoded by the lectin Yos9, an ERAD substrate recognition factor (Quan et al., 2008; Clerc et al., 2009).Open in a separate windowFigure 2.Mechanisms of substrate recognition in ERAD. (A) Recognition of misfolded luminal glycoproteins in yeast. Newly synthesized glycoproteins are bound by lectins and other chaperones which facilitate their folding. If properly folded, the proteins leave the ER. Prolonged residency in the ER, indicative of a persistent misfolded domain (red star), leads to Htm1-dependent exposure of an α-1,6–linked mannose residue (red bar). Together, the misfolded domain and the terminal α-1,6 mannose form the degradation signal recognized by Hrd3/Yos9. (B) Recognition of native MHC I heavy chain by the cytomegalovirus -encoded US2 adaptor in infected cells. US2 binds to folded MHC I in the ER membrane and delivers it to an E3 ligase complex containing the E3 Trc8 and the signal-peptide peptidase SPP resulting in MHC I degradation by ERAD. (C) Sterol-dependent recognition of HMGR by Insigs in mammalian cells. Under low sterol levels, HMGR is a stable protein at the ER membrane. High sterol levels, in particular the accumulation of 24,25-dihydrolanosterol (four-ringed structure in gray), cause Insig to bind to HMGR and to deliver it to an E3 ligase complex that promotes HMGR degradation by ERAD. The p97 cofactors Npl4 and Ufd1 are depicted as N and U, respectively. Ubiquitin is depicted as small circles.Both yeast Htm1 and its mammalian counterpart EDEM are in complex with oxidoreductases (Pdi1 in yeast, Erdj5 in mammals), required for the stability of Htm1 and also for reducing disulfide bonds in misfolded proteins, which might affect subsequent ERAD steps (Ushioda et al., 2008; Clerc et al., 2009). The binding of Yos9 to the α1,6-linked mannose is not sufficient to trigger the degradation of the misfolded protein. This processed glycan must be located in an unstructured polypeptide segment that is bound by Hrd3 (Xie et al., 2009). The delivery of substrates to Hrd3 might be facilitated by the luminal chaperone Kar2 that is essential for ERAD (Plemper et al., 1997; Denic et al., 2006; Xie et al., 2009). The dual recognition of a specific N-linked glycan by Yos9 and an unstructured segment by Hrd3 likely enhances the stringency of ERAD substrate recognition, perhaps by a kinetic proofreading mechanism (Fig. 2 A; Denic et al., 2006).The recognition mechanism of misfolded luminal N-linked glycoproteins is likely similar in mammalian cells because the yeast components are largely conserved in higher eukaryotes (Christianson et al., 2008). Whether the interactions with Sel1 and the substrates occur sequentially or correspond to different pools of OS-9/XTP3-B is unclear. In either case, using the same domain to interact with a component and a substrate offers OS-9/XTP3-B an additional mechanism to regulate recognition of N-glycosylated ERAD substrates.Recent work in yeast suggests that a different type of glycosylation, O-mannosylation, plays an important role in removing certain luminal proteins from futile folding cycles and thus favoring their degradation by ERAD after prolonged residency in the ER (Goder and Melero, 2011; Xu et al., 2013). The enzymes involved in protein O-mannosylation physically associate with ERAD machinery, but how O-mannosylated proteins are captured by the ERAD components is still unclear (Goder and Melero, 2011). Therefore, O-mannosylation is another appealing mechanism for generating an irreversible biochemical mark on proteins displaying folding problems.A common feature between ERAD substrate recognition by N-glycan trimming and O-mannosylating enzymes is that both appear to be slow processes, requiring substrates to stay for relatively prolonged periods in the ER. Whether other mechanisms involved in recognition of misfolded proteins by ERAD also require a lag period in the ER is not known. Nevertheless, it is curious that newly synthesized proteins were shown to be protected from degradation for a period of time, even under conditions that favor their misfolding (Vabulas and Hartl, 2005). Therefore, recognition of misfolded proteins might have evolved as an intrinsically slow process, perhaps to spare some folding intermediates from prematurely engaging in ERAD.

Adaptor-mediated substrate recognition.

The degradation of specific folded proteins by ERAD is mediated by the same general machinery, but the recognition of these substrates involves distinct factors. A simple mechanism to target a native protein to ERAD is by a substrate-specific adaptor. For example, the human cytomegalovirus encodes ER membrane adaptor proteins, US2 and US11, which bind independently to newly synthesized MHC I molecules and deliver them to ERAD components for degradation. As a consequence, infected cells display less MHC I complex at their surface and escape detection by the immune system (Wiertz et al., 1996a). Despite this common outcome, the two viral proteins interact differently with ERAD components. US11 uses its transmembrane domain to recruit MHC I into a complex which contains Derlin-1 as well as Sel1L, the p97 ATPase complex and its membrane adaptor UBXD8 (Lilley and Ploegh, 2004; Ye et al., 2004; Mueller et al., 2008). Intriguingly, the E3 ligase required for US11-mediated MHC I degradation is not known and both Hrd1 and Gp78 E3 ligases, which are found in complex with Derlin-1, appear to be dispensable. US2, on the other hand, delivers its substrate to the ERAD ligase complex containing the E3 Trc8 and SPP, the signal peptide peptidase (Fig. 2 B; Loureiro et al., 2006; Stagg et al., 2009). The precise function of SPP in ERAD is not known and it is not even clear whether it involves its proteolytic activity. Despite these differences, both US2 and US11 act as adaptors to deliver a specific ERAD substrate, MHC I heavy chain, to the E3 ligase complexes that promote its degradation.A similar mechanism targets CD4 for degradation in cells expressing the HIV-encoded ER membrane protein Vpu. In this case, Vpu works not only as the substrate adaptor for CD4 but also as a scaffold to recruit a cytosolic E3 ligase complex, SCFβTrCP, required for CD4 ubiquitination (Fujita et al., 1997; Schubert et al., 1998). In vitro reconstitution of Vpu-mediated CD4 ubiquitination revealed that the specificity of adaptor-mediated substrate selection can be further increased at the level of substrate ubiquitination, which can be counteracted by the activity of de-ubiquitinating enzymes (Zhang et al., 2013). The balance between these activities helps discriminating small differences in adaptor–substrate affinity. Whether this mechanism aids the selection of other ERAD substrates is not known yet.The strategy of using a substrate-specific adaptor is not exclusive to viral encoded proteins. Both in Drosophila and in mammalian cells, the derlin-related iRhom proteins function as adaptors in the ERAD-mediated degradation of EGFR ligands as they traffic through the ER (Zettl et al., 2011). Elegant genetic experiments in flies showed that this mode of regulated ERAD was important to control sleeping behavior that requires EGFR signaling (Zettl et al., 2011). It is likely that more substrate-specific adaptors will be identified as our knowledge of the mechanisms of regulated ERAD expands.

Signal-mediated substrate recognition.

A substrate-specific adaptor also functions in the regulated ERAD of HMGR, a key enzyme of the sterol biosynthetic pathway. In this case the adaptor, either Insig-1 or Insig-2, does not bind constitutively to HMGR (Song et al., 2005). Instead, the interaction only occurs in the presence of 24,25-dihydrolanosterol, an intermediate metabolite in sterol biosynthesis. Under low sterols levels HMGR is a stable protein, actively producing sterol precursors (Fig. 2 C). On the other hand, high sterol synthesis leads to a rise in 24,25-dihydrolanosterol concentration, which favors the binding of HMGR to one of the Insig proteins, its delivery to an E3 ligase complex, and consequently its degradation by the proteasome (Fig. 2 C). Whereas Gp78 and the Trc8 were originally implicated in HMGR regulated ERAD, recent data suggest that additional E3 ligases might also be involved (Song et al., 2005; Lee et al., 2010; Jo et al., 2011; Tsai et al., 2012). Degradation of HMGR by ERAD results in reduced flux through the sterol biosynthetic pathway and reestablishment of membrane lipid homeostasis.Interestingly, Insig-1 (but not Insig-2) is itself subjected to reciprocal sterol-regulated ERAD (Lee et al., 2006). Depletion of cellular sterols stimulates Insig-1 ubiquitination by the E3 Gp78 ligase complex. Conversely, if sterol levels are high Insig-1 binds to SCAP, another key ER membrane protein required for sterol homeostasis, leading to a much longer Insig-1 half-life. These data illustrate the complex interplay between the opposing effects of sterols on the stability of the HMGR enzyme and one of its adaptors for regulated degradation by ERAD.In yeast, sterol homeostasis also involves negative feedback of an HMGR homologue, Hmg2 (Hampton et al., 1996). Like in mammals, degradation of yeast Hmg2 is controlled by the Insig-like proteins Nsg1 and Nsg2 and requires the E3 ligase Hrd1 (Bays et al., 2001a; Gardner et al., 2001; Flury et al., 2005). In fact, Hrd1 was originally identified in a genetic screen for mutants defective in HMGR degradation (HRD genes; Hampton et al., 1996). The binding of Hmg2 to Nsg1 is also modulated by sterol levels (Theesfeld and Hampton, 2013). However, in contrast to mammalian cells, the binding of Nsg1 promotes Hmg2 stability, indicating that the recognition of this substrate is mechanistically different in the two systems (Flury et al., 2005; Theesfeld and Hampton, 2013). Based on limited proteolysis experiments, it has been proposed that Nsg1 and Nsg2 work as Hmg2-specific chaperones and that in their absence Hmg2 presents sufficient conformation instability to engage in ERAD as a misfolded protein (Flury et al., 2005; Shearer and Hampton, 2005). This degradation is further accelerated by high concentrations of an early sterol-intermediate metabolite, geranylgeranyl pyrophosphate (Theesfeld and Hampton, 2013). Therefore, a change in the affinity to a binding partner is another strategy to target a protein for ERAD in a signal-dependent manner.Squalene monooxygenase (SQLE), another enzyme of the sterol biosynthetic pathway, is also targeted by regulated ERAD both in yeast and in mammals (Foresti et al., 2013). SQLE degradation requires the yeast Doa10 E3 ligase complex or its mammalian homologue Teb4, indicating that two independent branches of ERAD control distinct steps in sterol biosynthesis (Foresti et al., 2013). Although the mechanism for the recognition of SQLE by ERAD machinery is still not known, it is clear that Insigs are dispensable for this recognition, both in mammals and in yeast (Gill et al., 2011; Foresti et al., 2013). In mammals, the N-terminal domain of SQLE is necessary and sufficient for cholesterol-dependent degradation (Gill et al., 2011). Whether this domain binds directly to cholesterol or interacts with an ERAD-specific adaptor is not known. The mechanism for SQLE recognition by ERAD in yeast is likely to be different because this N-terminal domain is only conserved among certain animals.Based on these few examples, it is clear that signal-mediated ERAD depends on the ability of cells to sense the concentration of some lipids in their membranes and on specific adaptors to selectively degrade key enzymes. Our knowledge on the mechanisms by which other classes of regulated ERAD substrates are recognized will grow as more of these are identified.

Shipping out the trash: Substrate retrotranslocation and cytoplasmic events in ERAD

After being selected, ERAD substrates are retrotranslocated across the ER membrane back into the cytoplasm. In the case of misfolded luminal proteins the complete polypeptide needs to be retrotranslocated, whereas for membrane substrates this step requires the transport of only certain domains. As a consequence, ubiquitination of luminal substrates only occurs at late stages of retrotranslocation, once a portion of the substrate has been exposed to the cytoplasm. In contrast, ubiquitination of most, but not all (Burr et al., 2013), membrane substrates is coupled to their retrotranslocation.In analogy to the transport of newly synthesized proteins into the ER or mitochondria, it has been postulated that retrotranslocation occurs through a protein-conducting channel. However, the identity of the retrotranslocation channel has been at the center of an intense debate that is almost as old as the research in this field. Over the years, several channel candidates have been proposed but it has been difficult to gather definitive evidence in support of any of them. The Sec61 translocon used for protein import into the ER was the first proposed retrotranslocation channel (Wiertz et al., 1996b). Sec61 was found to interact with ERAD substrates both in yeast and in mammalian cells as well as with the yeast proteasome (Wiertz et al., 1996b; Kalies et al., 2005; Scott and Schekman, 2008). However, the significance of these associations is not clear. Recent work showed that proteins engaging the Sec61 translocon aberrantly or persistently in their way into the ER become substrates of the Hrd1 ligase complex, which might explain the interaction between Sec61 and some ERAD substrates (Rubenstein et al., 2012). In addition, certain yeast sec61 mutants displayed defects in degrading model ERAD substrates, even under conditions in which general “forward” translocation appeared not to be dramatically affected (Pilon et al., 1997; Plemper et al., 1997; Willer et al., 2008). It remains to be determined whether this phenotype is caused by a specific impairment in retrotranslocation.The E3 ligase complexes interact with ERAD substrates immediately before and after their retrotranslocation, indicating this step occurs in their immediate vicinity. Therefore, multispanning membrane proteins within the E3 ligase complexes have also been seen as good candidates to mediate retrotranslocation (Ye et al., 2001; Lilley and Ploegh, 2004; Kreft et al., 2006; Horn et al., 2009; Carvalho et al., 2010; Mehnert et al., 2014). These include the E3 ligases themselves as well as proteins of the Derlin family (Der1 in yeast), which are essential for the degradation of all luminal ERAD substrates but whose function has remained elusive. In vitro experiments using mammalian-derived microsomes loaded with the yeast ERAD substrate mutant pro-α-factor indicate that Derlin might be involved in retrotranslocation (Wahlman et al., 2007). In this simplified system, in which synthesis and retrotranslocation were uncoupled, the substrate cross-linked to Derlin but not to Sec61. Moreover, its retrotranslocation was blocked by anti-Derlin antibodies whereas antibodies directed to Sec61 had no effect (Wahlman et al., 2007). It should be noted that mutant pro-α-factor is a noncanonical substrate because its retrotranslocation does not require ubiquitination. Interactions between the yeast Der1 and the prototype ERAD-L substrate CPY* were also detected by site-specific photocrosslinking in yeast (Carvalho et al., 2010; Stanley et al., 2011; Mehnert et al., 2014). These cross-links were seen even in cells lacking the substrate recognition factors Hrd3 and Yos9, and were increased if conserved polar residues in the membrane domain of Der1 were mutated. An interpretation of these results is that Der1 mediates the transfer of substrates from the recognition factors Hrd3/Yos9 to the Hrd1 ligase inside the membrane (Mehnert et al., 2014).A compelling piece of evidence for a function during substrate retrotranslocation was reported for the E3 ligase Hrd1 (Fig. 3; Carvalho et al., 2010). Although commonly working in the context of the Hrd1 complex, overexpressed Hrd1 can mediate the degradation of ERAD-L substrates even in the absence of its membrane partners Hrd3, Der1, and Usa1 (Plemper et al., 1999; Denic et al., 2006; Carvalho et al., 2010). Under these conditions Hrd1 selectivity for misfolded proteins is lost, suggesting that the other subunits of the complex are critical to control Hrd1 activity and substrate specificity (Denic et al., 2006). Hrd1 interacts with a sizeable region of a modified version of CPY* during the early stages of retrotranslocation, as assayed by site-specific photocrosslinking (Carvalho et al., 2010). Importantly, the interaction likely occurs inside of the ER bilayer because it is lost if substrate recognition is blocked or if the transmembrane segments of Hrd1 are mutated. Hrd1 contains only six transmembrane segments; therefore, ERAD-L substrate retrotranslocation requires Hrd1 oligomerization, which normally is facilitated by Usa1 but can occur spontaneously upon Hrd1 overexpression (Horn et al., 2009; Carvalho et al., 2010). All these data make a strong case for a direct role of Hrd1 in the retrotranslocation of ERAD-L substrates, but the possibility that it works with some partner(s), such as Der1, Sec61, or other unknown factors cannot be excluded at this point. Moreover, it is not known whether this is a unique feature of Hrd1 or whether the membrane domains of other E3 ligases also participate in the retrotranslocation of other classes of ERAD substrates.Open in a separate windowFigure 3.A working model for Hrd1-mediated retrotranslocation of a luminal misfolded glycoprotein. Upon recognition (not depicted), the misfolded protein (gray) is transferred to Hrd1. The binding can occur either to Hrd1 monomers or to Usa1-mediated Hrd1 dimers (1). Substrate-bound Hrd1 dimer self-ubiquitinates (2), which leads to the recruitment of the Cdc48 ATPase complex. ATP hydrolysis by Cdc48 induces a conformational change in Hrd1 dimer that facilitates the early stages of substrate retrotranslocation (3). Once exposed to the cytoplasm, the substrate is ubiquitinated by Hrd1 and recognized by the Cdc48 complex (4), which uses its ATPase activity to complete substrate retrotranslocation (5). After retrotranslocation, the ubiquitinated substrate is released in the cytosol for degradation by the proteasome (6). The Cdc48 cofactors Npl4 and Ufd1 are depicted as N and U, respectively. Ubiquitin is depicted as small circles.In most of these cross-linking experiments the retrotranslocation of CPY* was dampened by fusing it to a very tightly folded domain, indicating that ERAD-L substrates need to be unfolded before this step (Bhamidipati et al., 2005; Carvalho et al., 2010). Whether unfolding is also a prerequisite for the retrotranslocation of other classes of ERAD substrates is not settled yet (Fiebiger et al., 2002; Tirosh et al., 2003).At the cytoplasmic side of the ER membrane, substrates are ubiquitinated, a modification that allows their recognition by the Cdc48/p97 ATPase complex composed of a homohexamer of Cdc48/p97 and by the cofactors Ufd1 and Npl4 (Bays et al., 2001b; Ye et al., 2001, 2003; Jarosch et al., 2002; Rabinovich et al., 2002). The recruitment of this ATPase complex to the ER membrane is facilitated by ubiquitin regulatory X (UBX) domain–containing proteins, Ubx2 in yeast and UBXD8 in mammals (Neuber et al., 2005; Schuberth and Buchberger, 2005; Mueller et al., 2008). The ATPase activity of the Cdc48/p97 complex provides the driving force to move ubiquitinated substrates out of the membrane into the cytosol (Ye et al., 2003). Although the role of Cdc48/p97 in this process is well established, the mechanism that couples ATP hydrolysis to membrane extraction of the substrate is still not understood. In addition to the Cdc48/p97 complex, the ATPase subunits of the proteasome regulatory particle were also shown to play a role in retrotranslocation of some ERAD substrates (Lipson et al., 2008). The driving force for the retrotranslocation of the few non-ubiquitinated substrates, like the cholera toxin or pro-α-factor, is not known (Kothe et al., 2005; Moore et al., 2013).The Cdc48/p97 complex also serves as a platform for other ubiquitin-modifying enzymes such as de-ubiquitinating enzymes (DUBs; Rumpf and Jentsch, 2006; Jentsch and Rumpf, 2007; Ernst et al., 2009). Although the role of some of these Cdc48/p97-binding factors is not clear yet, it was shown that interfering with the DUBs YOD1 and ataxin-3 affected substrate retrotranslocation (Wang et al., 2006; Ernst et al., 2009). The requirement for DUB activity during retrotranslocation suggests that the process involves cycles of ubiquitination and de-ubiquitination. In some cases these cycles might be important to increase the specificity of substrate recognition, as was shown for the Vpu-mediated degradation of CD4 (Zhang et al., 2013). Interestingly, the retrotranslocation of some noncanonical substrates like cholera toxin, which are not ubiquitinated, is also affected by manipulation of both E3 ligase and DUB activities (Hassink et al., 2006; Bernardi et al., 2013). These results suggest that retrotranslocation requires the ubiquitination of some factors other than the substrates. Future studies should test some obvious candidates such as the components of the E3 ligase complexes themselves.There is some recent evidence that the yeast Cdc48 complex might be acting also much earlier, during the initial stages of retrotranslocation of an ERAD-L substrate, before it is exposed to the cytoplasm (Fig. 3). Using a cross-linking strategy, it was shown that the interaction between Hrd1 and an early retrotranslocation intermediate was lost in mutants of the Cdc48 complex (Carvalho et al., 2010). Interestingly, a similar defect was detected in Hrd1 mutants impaired for E3 ligase activity (Carvalho et al., 2010). Based on these observations it was proposed that in early stages of ERAD-L substrate retrotranslocation, Hrd1 induces the ubiquitination of an ERAD component, perhaps Hrd1 itself. This ubiquitination signals the recruitment of the Cdc48 complex, which upon ATP hydrolysis would induce changes in the conformation or in the oligomeric status of Hrd1, resulting in substrate retrotranslocation. This model is consistent with the well-established role of Cdc48/p97 in the disassembly and remodeling of protein complexes (Jentsch and Rumpf, 2007). Once the substrate emerges on the cytosolic side and is ubiquitinated by Hrd1, the ATPase complex binds to it and promotes the final stages of its retrotranslocation. Although many aspects of this model still wait for experimental support, it would provide a unifying role for the Cdc48/p97 ATPase as the driving force for substrate retrotranslocation in ERAD (Fig. 3).After being released from the membrane, substrates are kept soluble and transferred to the proteasome by cytosolic chaperones such as the BAG6 complex (Claessen and Ploegh, 2011; Wang et al., 2011; Xu et al., 2012) or shuttle factors like Rad23 and Dsk2 (Medicherla et al., 2004). The long journey of ERAD substrates ends with their degradation by the proteasome.

Conclusions and future perspectives

In recent years there has been tremendous progress in understanding ERAD. The identification of most of the components involved in this process and how these are pieced together and organized in the different ERAD branches were important achievements. A major challenge for the future is to reveal the mechanistic aspects of the pathway. Such developments should, for example, help in discerning the basis by which misfolded proteins are recognized in each of the different ERAD branches.The mechanisms of substrate retrotranslocation and the roles played by the different components, such as the E3 ligases and the Cdc48/p97 complex, will certainly be another interesting area to follow. However, progress on these topics might require the development of new approaches, such as in vitro systems with purified components recapitulating individual ERAD steps.In early days, ERAD was perceived as a process mainly dedicated to ER protein quality control. The picture now emerging places ERAD closer to other ubiquitination systems in the cytoplasm and nucleus, which control the turnover of specific proteins to achieve a certain physiological state. Therefore, another major challenge for the coming years is the detailed characterization of the roles of ERAD beyond quality control. Although the central role of ERAD in sterol homeostasis is unequivocal, it will be important to clarify whether ERAD has a more general role in the regulated degradation of folded ER proteins and in that way modulates other ER-related functions. A systematic and rigorous identification of regulated ERAD substrates should help in addressing these issues. Uncovering the intersections of ERAD with other cellular pathways will provide important insights into the mechanisms of ER and cellular homeostasis.  相似文献   

4.
5.
6.
The concept that target tissues determine the survival of neurons has inspired much of the thinking on neuronal development in vertebrates, not least because it is supported by decades of research on nerve growth factor (NGF) in the peripheral nervous system (PNS). Recent discoveries now help to understand why only some developing neurons selectively depend on NGF. They also indicate that the survival of most neurons in the central nervous system (CNS) is not simply regulated by single growth factors like in the PNS. Additionally, components of the cell death machinery have begun to be recognized as regulators of selective axonal degeneration and synaptic function, thus playing a critical role in wiring up the nervous system.

Why do so many neurons die during development?

Programmed cell death occurs throughout life, as cell turnover is part of homeostasis and maintenance in most organs and tissues. The situation in the nervous system is principally different, as the vast majority of neurons undergo their last round of cell division early in development. Soon after exiting the cell cycle, neurons start elongating axons to innervate their targets. It is during this period that they are highly susceptible to undergo programmed cell death: a large percentage, as much as 50% in several ganglia in the peripheral nervous system (PNS) as well as in various central nervous system (CNS) areas, is eliminated around the time that connections are being made with other cells. Later in development, the propensity of neurons to initiate apoptosis progressively decreases. The likelihood for a neuron to undergo apoptosis seems to be determined by a tightly regulated apoptotic machinery (summarized in Fig. 1). Therefore, modulation of the expression levels or the activity of components of this apoptotic balance changes the sensitivity to death-promoting cues, allowing temporal restriction of cell death.Open in a separate windowFigure 1.Core components of the apoptotic machinery. The likelihood that a neuron undergoes apoptosis is determined by the interplay of the tightly interlinked apoptotic machinery, many components of which are highly conserved between species. The critical, and often terminal, step in programmed cell death is the proteolytic activation of the executor caspases (such as caspase 3, 6, 7) by the initiator caspases (i.e., caspase 8, 9, and 10; Riedl and Salvesen, 2007). In mammalian cells, initiation of the executor caspases is regulated by two distinct protein cascades: the intrinsic pathway, also known as the mitochondrial pathway, and the extrinsic pathway. The intrinsic pathway integrates a number of intra- and extracellular signal modalities, such as redox state (for example, the reactive oxygen species; Franklin, 2011), DNA damage (Sperka et al., 2012), ER stress (Puthalakath et al., 2007) and growth factor deprivation (Deckwerth et al., 1998; Putcha et al., 2003; Bredesen et al., 2005), or activation of the p75NTR neurotrophin receptor by pro-neurotrophins (Nykjaer et al., 2005). The stressors converge onto pro- and anti-apoptotic members of the Bcl-2 protein family (for example: BCL-2, BCL-Xl, BAX, and tBID; Youle and Strasser, 2008). These proteins regulate the release of cytochrome c from mitochondria, which activates the initiator caspase 9 through Apaf1 (Riedl and Salvesen, 2007). The extrinsic pathway links activation of ligand-bound death receptors (such as Fas/CD95 and TNFR) to the initiator caspase 8 and 10, through formation of the death-inducing signaling complex (DISC; LeBlanc and Ashkenazi, 2003; Peter and Krammer, 2003). Together with additional regulatory elements (including the Inhibitors of apoptosis proteins [IAP]; Vaux and Silke, 2005) and cFLIP (Scaffidi et al., 1999; Wang et al., 2005), the apoptotic machinery forms a balance that determines the propensity of the neuron to undergo apoptosis.Programmed cell death eliminates many neurons during development, even in organisms comprised of only few cells, such as Caenorhabditis elegans. As neurons and their targets are initially separated, it is possible that the initial generation of an overabundance of neurons is simply part of a mechanism to ensure that distal targets are adequately innervated (Oppenheim, 1991; Conradt, 2009; Chen et al., 2013). In various tissues other than the nervous system, programmed cell death is used to eliminate cells that are no longer needed, defective, or harmful to the function of the organism. However, there is strong evidence that the elimination of superfluous neurons in the developing nervous system is not essential. For example, early work in C. elegans revealed that preventing programmed cell death does not result in significant behavioral alterations (Ellis and Horvitz, 1986). In the C57BL/6 mouse strain, deletion of the executor caspases 3 and 7 (Fig. 1) has a remarkably limited neuropathological and morphological impact in the CNS (Leonard et al., 2002; Lakhani et al., 2006) compared with the 129X1/SvJ strain, in which deletion of these caspases causes neurodevelopmental defects (Leonard et al., 2002). Similar conclusions were reached by blocking the Bcl-2–associated X protein (BAX)–dependent pathway in many neuronal populations, including motoneurons (Buss et al., 2006a). A recent study in the developing retina showed that in mice lacking the central apoptotic regulator BAX, the normal mosaic distribution of intrinsically photosensitive retinal ganglion cells (ipRGCs) was perturbed (Chen et al., 2013). Although this abnormal distribution is dispensable for the intrinsic photosensitivity of the ipRGCs, it is required for establishing proper connections to other neurons in the retina, which is necessary for rod/cone photo-entrainment (Chen et al., 2013). Even though this finding highlights a physiological role for programmed cell death in the CNS, the functional consequences remain rather underwhelming in the face of a process that eliminates such large numbers of neurons (Purves and Lichtman, 1984; Oppenheim, 1991; Miller, 1995; Gohlke et al., 2004). It thus appears that apoptotic removal of the surplus neurons generated during development mainly serves the purpose to optimize the size of the nervous system to be minimal, but sufficient.

A molecular substrate for the neurotrophic theory

Quantitatively, programmed cell death of neurons in the PNS and CNS is most dramatic when neurons start contacting the cells they innervate. Because experimental manipulations such as target excision typically lead to the death of essentially all innervating neurons (Oppenheim, 1991), the concept emerged that the fate of developing neurons is regulated by their targets. This notion is often referred to as the “neurotrophic theory” (Hamburger et al., 1981; Purves and Lichtman, 1984; Oppenheim, 1991), but it is important to realize that it evolved in the absence of direct mechanistic or molecular support (Purves, 1988). Originally described as a diffusible agent promoting nerve growth, the eponymous NGF later provided a strong and very appealing molecular foundation for this theory (Korsching and Thoenen, 1983; Edwards et al., 1989; Hamburger, 1992). The tyrosine kinase receptor tropomyosin receptor kinase A (TrkA), which was initially identified as an oncogene (Martin-Zanca et al., 1986), was fortuitously discovered to be the critical receptor necessary for the prevention of neuronal cell death by NGF (Klein et al., 1991). Both the remarkable expression pattern of TrkA in NGF-dependent neurons and the onset of its expression during development (Martin-Zanca et al., 1990) provided further additional support for the neurotrophic theory. However, for a surprisingly long time, the question was not asked as to why only specific populations of neurons strictly depend on NGF for survival, while others do not. Indeed, it was only recently shown that TrkA causes cell death of neurons by virtue of its mere expression, and that this death-inducing activity is prevented by addition of NGF (Nikoletopoulou et al., 2010). These findings thus indicate that TrkA acts as a “dependence receptor,” a concept introduced after observations that various cell types die when receptors are expressed in the absence of their cognate ligands (Bredesen et al., 2005; Tauszig-Delamasure et al., 2007). Accordingly, embryonic mouse sympathetic or sensory neurons survive in the absence of NGF when TrkA is deleted (Nikoletopoulou et al., 2010). The closely related neurotrophin receptor TrkC also acts as a dependence receptor (Tauszig-Delamasure et al., 2007; Nikoletopoulou et al., 2010). Here, it is interesting to note a series of older, convergent results indicating that deletion of neurotrophin-3 (NT3), the TrkC ligand, leads to a significantly larger loss of sensory and sympathetic neurons in the PNS than the deletion of TrkC (Tessarollo et al., 1997). This phenotypic discrepancy fits well with the idea that inactivation of the ligand of a dependence receptor is expected to yield a more profound phenotype than inactivation of the receptor itself (Tauszig-Delamasure et al., 2007). How TrkA and TrkC induce apoptosis remains to be fully elucidated. It seems that proteolysis is involved, either of TrkC itself (Tauszig-Delamasure et al., 2007), as was suggested for other dependence receptors (Bredesen et al., 2005), or by the proteolysis of the neurotrophin receptor p75NTR, which associates with both TrkA and TrkC (Fig. 2; Nikoletopoulou et al., 2010). Surprisingly, although TrkA and TrkC cause cell death, the structurally related TrkB receptor does not (Nikoletopoulou et al., 2010), a difference that appears to be accounted for by their differential localization in the cell membrane. TrkA and TrkC colocalize with p75NTR in lipid rafts, whereas TrkB, which also associates with p75NTR (Bibel et al., 1999), is excluded from lipid rafts (Fig. 2; unpublished data). Interestingly, the transmembrane domains of TrkA and TrkC are closely related, and differ clearly from that of TrkB. It turns out that a chimeric protein of TrkB with the transmembrane domain of TrkA causes cell death, which can be prevented by the addition of the TrkB ligand brain-derived neurotrophic factor (BDNF; unpublished data). The suggestion that the lipid raft localization of TrkA and TrkC is important for their death-inducing function is in line with a number of reports indicating that certain apoptotic proteins preferentially localize in lipid rafts in the plasma membrane. After activation of the extrinsic apoptosis pathway, translocation of the activated receptors to lipid rafts in the membrane is required for assembling the death-inducing signaling complex (DISC; Davis et al., 2007; Song et al., 2007). Indeed, regulators of the extrinsic pathway (e.g., cFLIP; Fig. 1) prevent this translocation, explaining how they attenuate cell death induction (Song et al., 2007). Similarly, the localization of the dependence receptor DCC (deleted in colorectal cancer) in lipid rafts is a prerequisite for its pro-apoptotic activity in absence of its ligand, Netrin-1 (Furne et al., 2006).Open in a separate windowFigure 2.TrkA and TrkC as dependence receptors: mode of action and contrast with TrkB. All Trk receptors associate with the pan-neurotrophin receptor p75NTR (Bibel et al., 1999). A critical step in the induction of apoptosis by TrkA is the release of the intracellular death domain of p75NTR by the protease γ-secretase (Nikoletopoulou et al., 2010), which is localized in lipid rafts (Urano et al., 2005). Our membrane fractionation studies indicate that while TrkA and TrkC associate with p75NTR in lipid rafts, TrkB associated with p75NTR is excluded from this membrane domain (unpublished data). The 24–amino acid transmembrane domain of the Trk receptors may be responsible for this differential localization (see text).Despite the fact that TrkB does not act as a dependence receptor, its activation by BDNF is required for the survival of several populations of cranial sensory neurons (Ernfors et al., 1995; Liu et al., 1995). It appears that other death-inducing receptors predispose these neurons to be eliminated, such as p75NTR, which is expressed at high levels in some of these ganglia, or TrkC in vestibular neurons (Stenqvist et al., 2005). This latter case is of special interest, as NT3 is known not to be required for the survival of these neurons (Stenqvist et al., 2005). In addition to inducing apoptosis in the absence of their ligand, TrkA and TrkC have long been recognized to have a pro-survival function similar to TrkB, as can be inferred from the loss of specific populations of peripheral sensory neurons in mutants lacking these receptors (Klein et al., 1994; Smeyne et al., 1994).

Cell death in the CNS

Although TrkA is primarily expressed in peripheral sympathetic and sensory neurons, it is also found in a small population of cholinergic neurons in the basal forebrain (Sobreviela et al., 1994), a proportion of which requires NGF for survival (Hartikka and Hefti, 1988; Crowley et al., 1994; Müller et al., 2012). Selective deletion of TrkA was recently shown not to cause the death of these neurons (Sanchez-Ortiz et al., 2012). This supports the notion that TrkA acts as a dependence receptor for this small population of CNS neurons, like for peripheral sensory and sympathetic neurons. TrkA activation by NGF is essential for the maturation, projections, and function of these neurons (Sanchez-Ortiz et al., 2012), as was previously described for sensory neurons in the PNS as well (Patel et al., 2000).Whether or not receptors other than TrkA act as dependence receptors in the CNS is an important open question, particularly because TrkB, which is expressed highly by most CNS neurons, does not act as a dependence receptor (Nikoletopoulou et al., 2010). In retrospect, the structural similarities between TrkA and TrkB, just like those between NGF and BDNF (Barde, 1989), have substantially misled the field by suggesting that BDNF would act in the CNS like NGF in the PNS. Adding to the confusion were early findings showing that BDNF supports the growth of spinal cord motoneurons in vitro or in vivo after axotomy (Oppenheim et al., 1992; Sendtner et al., 1992; Yan et al., 1992). However, in the absence of lesion, deletion of BDNF does not lead to significant losses of neurons in the developing or adult CNS (Ernfors et al., 1994a; Jones et al., 1994; Rauskolb et al., 2010), unlike the case in some populations of PNS neurons. The poor correlation of the role of BDNF in CNS development and in axotomy and in vitro experiments is surprising, especially because the role of NGF in vivo could in essence be recapitulated by in vitro experiments. Although the reasons for this discrepancy are not fully understood, the strong up-regulation of death-inducing molecules such as p75NTR after axotomy (Ernfors et al., 1989) may be a part of the explanation. At present, most of the growth factors promoting the survival of PNS neurons fail to show significant survival properties for developing neurons in the CNS, as for example was shown for NT3 (Ernfors et al., 1994b; Fariñas et al., 1994), glial cell line–derived neurotrophic factor (GDNF; Henderson et al., 1994), ciliary neurotrophic factor (CNTF; DeChiara et al., 1995), and several others.In the developing CNS, neuronal activity and neurotransmitter input seem to play a more significant role than single growth factors in regulating neuronal survival. In particular, it has been known for a long time that blocking synaptic transmission at the neuromuscular junction has a pro-survival effect on spinal cord motoneurons (Pittman and Oppenheim, 1978; Oppenheim et al., 2008). By contrast, surgical denervation of afferent connections leads to increased apoptosis of postsynaptic neurons (Okado and Oppenheim, 1984), whereas inhibiting glycinergic and GABAergic synaptic transmission has both pro- and anti-apoptotic effects on motoneurons (Banks et al., 2005). Throughout the developing brain, blocking glutamate-mediated synaptic transmission involving NMDA receptors markedly increases normally occurring neuronal death (Ikonomidou et al., 1999; Heck et al., 2008). The mechanism involves a reduction of neuronal expression of anti-apoptotic proteins, such as B-cell lymphoma 2 (BCL-2; Hansen et al., 2004). Conversely, a limited increase in neuronal activity leads to down-regulation of the pro-apoptotic genes BAX and caspase 9 (Léveillé et al., 2010), thereby reducing the propensity of these cells to initiate programmed cell death (Hardingham et al., 2002). In addition to directly modulating the expression of apoptotic proteins, neuronal activity affects the expression of several secreted growth factors, such as BDNF (Hardingham et al., 2002; Hansen et al., 2004) and GDNF (Léveillé et al., 2010). So, even though BDNF is not a major survival factor in the developing CNS, it appears to be critical for activity-dependent neuroprotection (Tremblay et al., 1999). A recent publication revealed that certain populations of neurons in the CNS do not follow the predictions of the neurotrophic theory and showed that apoptosis of cortical inhibitory neurons is independent of cues present in the developing cerebral cortex (Southwell et al., 2012). This study indicates that programmed cell death of a large proportion of interneurons in the CNS is regulated by intrinsic mechanisms that are largely resistant to the presence or absence of extrinsic cues (Dekkers and Barde, 2013).Taken together, even though the extent of naturally occurring cell death in the different regions of the CNS is not nearly as well characterized as in the PNS, let alone quantified, it appears that its regulation may significantly differ. Although single secreted neurotrophic factors seem to be largely dispensable for survival, neuronal activity and other intrinsic mechanisms drive the propensity of the neurons in the CNS to undergo apoptosis. An important open question in this context is a possible involvement of non-neuronal cells, such as glial cells (see Corty and Freeman, in this issue).

The apoptotic machinery as a regulator of connectivity

Activation of the executor caspases has been most studied in cell bodies and typically results in the demise of the entire cell (Williams et al., 2006). However, recent evidence shows that caspases are also activated locally in neuronal processes and branches destined to be eliminated, for example in axons overshooting their targets that are subsequently pruned back to establish the precise adult connectivity (Finn et al., 2000; Raff et al., 2002; Luo and O’Leary, 2005; Buss et al., 2006b). Initially, axonal degeneration and axon pruning were thought to be independent of caspases (Finn et al., 2000; Raff et al., 2002). Later work in Drosophila melanogaster (Kuo et al., 2006; Williams et al., 2006) and in mammalian neurons (Plachta et al., 2007; Nikolaev et al., 2009; Vohra et al., 2010) demonstrated that interfering with the apoptotic balance or the executor caspases can prevent or at least delay axonal degeneration. Simon et al. (2012) have found that a caspase 9 to caspase 3 cascade is crucial for axonal degeneration induced by NGF withdrawal, with caspase 6 activation playing a significant but subsidiary role. Upstream of the caspases, BCL-2 family members such as BAX and BCL-Xl are required (Nikolaev et al., 2009; Vohra et al., 2010). It is conceivable that the failure of ipRGCs in BAX-deficient mice to form appropriate connections to other cells in the retina (Chen et al., 2013) may be in part attributable to defective axonal degeneration. Surprisingly, Apaf1 appears not to be involved in this process (Cusack et al., 2013), suggesting that axon degeneration depends on the concerted activation of the intrinsic initiator complex in a different way from apoptosis.Strikingly, a series of recent studies showed that several caspases and components of the intrinsic pathway also affect normal synaptic physiology in adulthood (Fig. 3, A–D). Here, pro-apoptotic proteins are predominantly involved in weakening the synapses, whereas the anti-apoptotic proteins have been mainly associated with synaptic strengthening (Fig. 3 B). In particular, caspase 3 promotes long-term depression (LTD), a stimulation paradigm that results in a period of decreased synaptic transmission (Li et al., 2010), and also prevents long-term potentiation (LTP), the converse situation leading to strengthened synaptic transmission (Jo et al., 2011). Likewise, the proapoptotic BCL-2 family members BAX and BAD stimulate LTD (Jiao and Li, 2011). By contrast, the anti-apoptotic protein BCL-Xl increases synapse numbers and strength (H. Li et al., 2008), and the inhibitor of apoptosis protein (IAP) family member survivin was reported to be involved in LTP in the hippocampus (Iscru et al., 2013) and in activity-dependent gene regulation (O’Riordan et al., 2008).Open in a separate windowFigure 3.Canonical and noncanonical functions of the apoptotic machinery. (A) The apoptotic machinery is not only involved in eliminating cells destined to die, but is also a central player in refining neuronal connectivity, by regulating synaptic transmission and by generating the adult connectivity through axon pruning (Luo and O’Leary, 2005; Hyman and Yuan, 2012). But how the canonical and noncanonical roles of the apoptotic machinery are interlinked and spatially restricted is not well understood. (B) In the adult nervous system, the pro-apoptotic proteins BAX, caspase 9, and caspase 3 promote weakening of synapses (long-term depression [LTD]; Li et al., 2010; Jiao and Li, 2011; Jo et al., 2011), while the anti-apoptotic proteins Bcl-Xl and the IAP survivin promote synaptic strengthening (long-term potentiation [LTP]; Li et al., 2008a; Iscru et al., 2013). It is unclear how the activation of these pathways is restricted to a single synapse, but a recent review suggested that the proteasomal degradation of activated caspases may prevent their diffusion (Hyman and Yuan, 2012). (C) Caspase activation is now known to be required for axon pruning during development to generate the adult refined connectivity (Luo and O’Leary, 2005; Simon et al., 2012). Different pathways are activated depending on the stimulus leading to degeneration. Growth factor deprivation during development leads to activation the executor caspases 3 and 6 (Simon et al., 2012) through the intrinsic apoptotic pathway, although its core protein Apaf1 does not seem to be required for this process (Cusack et al., 2013). On the other hand, a traumatic injury leads to reduced influx of NMNAT2 into the axon, which negatively affects the stability and function of mitochondria and leads to an increased calcium concentration (Wang et al., 2012). The effector caspase, caspase 6, is dispensable for this form of axonal degeneration (Vohra et al., 2010; Simon et al., 2012). Regulatory proteins such as the IAPs and also the proteasome seem to play a role in limiting the extent of activation to the degenerating part of the axon (Wang et al., 2012; Cusack et al., 2013; Unsain et al., 2013). (D) Simplified schematic of the main pro- and anti-apoptotic components. DISC, death-induced signaling complex. IAP, inhibitor of apoptosis protein. See Fig. 1 for details.These findings indicate that the apoptotic machinery acts at different levels in the cell, ranging from driving sub-lethal degradation of a compartment (Fig. 3 C) and attenuating synaptic transmission at the neuronal network level (Fig. 3 B) to destroying the entire cell during development or in disease (Fig. 3 D). How the cell spatially restricts the extent of activation of the apoptotic machinery is yet unclear. For example, elimination of the somata of developing neurons after neurotrophin deprivation is preceded by axonal degeneration, but not all instances of axonal degeneration lead to the death of the neuron (Campenot, 1977; Raff et al., 2002). Local regulation of caspase activation by IAPs is well established as a means for ensuring the elimination of neuronal processes in D. melanogaster (Kuo et al., 2006; Williams et al., 2006). Recent findings suggest a similar role for IAP in mammalian neurons, where it limits caspase activation to the degenerating axon (Fig. 3 C; Cusack et al., 2013; Unsain et al., 2013). The spontaneous mutation Wallerian degeneration slow (WldS; Lunn et al., 1989) has been instrumental to understand that trauma-induced axon degeneration is a regulated process different from, and independent of, cell body degeneration (Wang et al., 2012), but also distinct from axon pruning (Hoopfer et al., 2006). Work on the chimeric protein encoded by the WldS mutation also led to the identification of the protein NMNAT2 (nicotinamide mononucleotide adenylyltransferase 2) as a labile axon survival factor (Gilley and Coleman, 2010). How the WldS chimeric protein and NMNAT2 result in axon protection is unclear, but several lines of evidence seem to converge on local regulation of mitochondrial function and motility (Avery et al., 2012; Fang et al., 2012).Related to the spatial limiting of apoptotic activity is the question of how a local source of neurotrophins leads to the rescue of a developing peripheral neuron. When neurons encounter a source of neurotrophins, only the receptors close to the target will be activated, whereas the others, located further away, are not. The cell, therefore, needs to integrate a pro-survival signal from the activated receptors, and death-inducing signals from the nonactivated dependence receptors. The continued signaling of activated neurotrophin receptors that are retrogradely transported to the soma (Grimes et al., 1996; Howe et al., 2001; Wu et al., 2001; Harrington et al., 2011) likely play a role in counteracting the pro-apoptotic signaling proximal to the source of neurotrophins. It will be interesting to investigate whether similar mechanisms play a role in axon pruning and traumatic axon degeneration as well.

Programmed cell death in the adult brain

Most of the nervous system becomes post-mitotic early in development. In rodents, two brain areas retain the capacity to generate new neurons in the adult: the sub-ventricular zone, which generates neurons that migrate toward the olfactory bulb, and the sub-granular zone of the dentate gyrus of the hippocampus, where neurons are generated that integrate locally. Similar to what is observed during embryonic development, these adult-generated neurons are produced in excess, and a large fraction undergoes apoptosis when contacting its designated targets (Petreanu and Alvarez-Buylla, 2002; Kempermann et al., 2003; Ninkovic et al., 2007). Preventing apoptosis of adult-generated neurons in the olfactory bulb only has limited functional consequences (Kim et al., 2007), whereas a similar maneuver in the dentate gyrus does lead to impaired performance in memory tasks (Kim et al., 2009). Why superfluous hippocampal neurons would need to be eliminated for proper function is a matter of speculation, but may be linked with the fact that these are excitatory projection neurons, whereas in the olfactory bulb only axon-less inhibitory granule cells are integrated. The extent of survival in both these areas critically depends on the activity of the neuronal network in which these newly born neurons have to integrate (Petreanu and Alvarez-Buylla, 2002; Kempermann et al., 2006; Ninkovic et al., 2007). In this context, BDNF, the expression level of which is well known to be regulated by network activity, supports the survival of young adult–generated neurons and possibly even stimulates the proliferation of neural progenitors (Y. Li et al., 2008; Waterhouse et al., 2012). Interestingly, in young adult mouse mutants that exhibit spontaneous epileptic seizures, significantly higher levels of BDNF have been measured (Lavebratt et al., 2006; Heyden et al., 2011). Concomitantly, the entire hippocampal formation is considerably enlarged by as much as 40% (Lavebratt et al., 2006; Angenstein et al., 2007), which in turn is dependent on the epileptic seizures (Lavebratt et al., 2006). Whether or not there is a causal relationship between increased BDNF levels and hippocampal volume remains to be established.

Conclusion

Now that is has become clear that action of the apoptotic machinery can be limited spatially and temporally, several questions need to be addressed: how do neurons integrate intrinsic and extrinsic pro- and anti-apoptotic signals; and how they are spatially restricted to allow degradation of a dendrite or axon, or modulation of synaptic transmission? Another important issue is the regulation of cell death by intrinsic mechanisms in the central nervous system of vertebrates, not least because programmed cell death is observed in the CNS in a number of neurodegenerative diseases (Vila and Przedborski, 2003). Indeed, several of the central apoptotic components discussed here are also involved in these disorders (Hyman and Yuan, 2012). New insights in the regulation of programmed cell death in the developing nervous system may therefore continue to help to better understand the pathophysiological mechanisms of neurodegenerative disorders.  相似文献   

7.
8.
9.
Microtubules are cytoskeletal filaments that are dynamically assembled from α/β-tubulin heterodimers. The primary sequence and structure of the tubulin proteins and, consequently, the properties and architecture of microtubules are highly conserved in eukaryotes. Despite this conservation, tubulin is subject to heterogeneity that is generated in two ways: by the expression of different tubulin isotypes and by posttranslational modifications (PTMs). Identifying the mechanisms that generate and control tubulin heterogeneity and how this heterogeneity affects microtubule function are long-standing goals in the field. Recent work on tubulin PTMs has shed light on how these modifications could contribute to a “tubulin code” that coordinates the complex functions of microtubules in cells.

Introduction

Microtubules are key elements of the eukaryotic cytoskeleton that dynamically assemble from heterodimers of α- and β-tubulin. The structure of microtubules, as well as the protein sequences of α- and β-tubulin, is highly conserved in evolution, and consequently, microtubules look alike in almost all species. Despite the high level of conservation, microtubules adapt to a large variety of cellular functions. This adaptation can be mediated by a large panel of microtubule-associated proteins (MAPs), including molecular motors, as well as by mechanisms that directly modify the microtubules, thus either changing their biophysical properties or attracting subsets of MAPs that convey specific functions to the modified microtubules. Two different mechanism can generate microtubule diversity: the expression of different α- and β-tubulin genes, referred to as tubulin isotypes, and the generation of posttranslational modifications (PTMs) on α- and β-tubulin (Figs. 1 and and2).2). Although known for several decades, deciphering how tubulin heterogeneity controls microtubule functions is still largely unchartered. This review summarizes the current advances in the field and discusses new concepts arising.Open in a separate windowFigure 1.Tubulin heterogeneity generated by PTMs. (A) Schematic representation of the distribution of different PTMs of tubulin on the α/β-tubulin dimer with respect to their position in the microtubule lattice. Acetylation (Ac), phosphorylation (P), and polyamination (Am) are found within the tubulin bodies that assemble into the microtubule lattice, whereas polyglutamylation, polyglycylation, detyrosination, and C-terminal deglutamylation take place within the C-terminal tubulin tails that project away from the lattice surface. The tubulin dimer represents TubA1A and TubB2B (Fig. 2), and modification sites for polyglutamylation and polyglycylation have been randomly chosen. (B) Chemical structure of the branched peptide formed by polyglutamylation and polyglycylation, using the γ-carboxyl groups of the modified glutamate residues as acceptor sites for the isopeptide bonds. Note that in the case of polyglutamylation, the elongation of the side chains generates classical peptide bonds (Redeker et al., 1991).Open in a separate windowFigure 2.Heterogeneity of C-terminal tails of tubulin isotypes and their PTMs. The amino acid sequences of all tubulin genes found in the human genome are indicated, starting at the last amino acid of the folded tubulin bodies. Amino acids are represented in single-letter codes and color coded according to their biochemical properties. Known sites for polyglutamylation are indicated (Eddé et al., 1990; Alexander et al., 1991; Rüdiger et al., 1992). Potential modification sites (all glutamate residues) are indicated. Known C-terminal truncation reactions of α/β-tubulin (tub) are indicated. The C-terminal tails of the yeast Saccharomyces cerevisiae are shown to illustrate the phylogenetic diversity of these domains.

Tubulin isotypes

The cloning of the first tubulin genes in the late 1970’s (Cleveland et al., 1978) revealed the existence of multiple genes coding for α- or β-tubulin (Ludueña and Banerjee, 2008) that generate subtle differences in their amino acid sequences, particularly in the C-terminal tails (Fig. 2). It was assumed that tubulin isotypes, as they were named, assemble into discrete microtubule species that carry out unique functions. This conclusion was reinforced by the observation that some isotypes are specifically expressed in specialized cells and tissues and that isotype expression changes during development (Lewis et al., 1985; Denoulet et al., 1986). These high expectations were mitigated by a subsequent study showing that all tubulin isotypes freely copolymerize into heterogeneous microtubules (Lewis et al., 1987). To date, only highly specialized microtubules, such as ciliary axonemes (Renthal et al., 1993; Raff et al., 2008), neuronal microtubules (Denoulet et al., 1986; Joshi and Cleveland, 1989), and microtubules of the marginal band of platelets (Wang et al., 1986; Schwer et al., 2001) are known to depend on some specific (β) tubulin isotypes, whereas the function of most other microtubules appears to be independent of their isotype composition.More recently, a large number of mutations in single tubulin isotypes have been linked to deleterious neurodevelopmental disorders (Keays et al., 2007; Fallet-Bianco et al., 2008; Tischfield et al., 2010; Cederquist et al., 2012; Niwa et al., 2013). Mutations of a single tubulin isotype could lead to an imbalance in the levels of tubulins as a result of a lack of incorporation of mutant isoforms into the microtubule lattice or to incorporation that perturbs the architecture or dynamics of the microtubules. The analysis of tubulin disease mutations is starting to reveal how subtle alterations of the microtubule cytoskeleton can lead to functional aberrations in cells and organisms and might provide novel insights into the roles of tubulin isotypes that have so far been considered redundant.

Tubulin PTMs

Tubulin is subject to a large range of PTMs (Fig. 1), from well-known ones, such as acetylation or phosphorylation, to others that have so far mostly been found on tubulin. Detyrosination/tyrosination, polyglutamylation, and polyglycylation, for instance, might have evolved to specifically regulate tubulin and microtubule functions, in particular in cilia and flagella, as their evolution is closely linked to these organelles. The strong link between those modifications and tubulin evolution has led to the perception that they are tubulin PTMs; however, apart from detyrosination/tyrosination, most of them have other substrates (Regnard et al., 2000; Xie et al., 2007; van Dijk et al., 2008; Rogowski et al., 2009).

Tubulin acetylation.

Tubulin acetylation was discovered on lysine 40 (K40; Fig. 1 A) of flagellar α-tubulin in Chlamydomonas reinhardtii (L’Hernault and Rosenbaum, 1985) and is generally enriched on stable microtubules in cells. Considering that K40 acetylation per se has no effect on the ultrastructure of microtubules (Howes et al., 2014), it is rather unlikely that it directly stabilizes microtubules. As a result of its localization at the inner face of microtubules (Soppina et al., 2012), K40 acetylation might rather affect the binding of microtubule inner proteins, a poorly characterized family of proteins (Nicastro et al., 2011; Linck et al., 2014). Functional experiments in cells have further suggested that K40 acetylation regulates intracellular transport by regulating the traffic of kinesin motors (Reed et al., 2006; Dompierre et al., 2007). These observations could so far not be confirmed by biophysical measurements in vitro (Walter et al., 2012; Kaul et al., 2014), suggesting that in cells, K40 acetylation might affect intracellular traffic by indirect mechanisms.Enzymes involved in K40 acetylation are HDAC6 (histone deacetylase family member 6; Hubbert et al., 2002) and Sirt2 (sirtuin type 2; North et al., 2003). Initial functional studies used overexpression, depletion, or chemical inhibition of these enzymes. These studies should be discussed with care, as both HDAC6 and Sirt2 deacetylate other substrates and have deacetylase-independent functions and chemical inhibition of HDAC6 is not entirely selective for this enzyme (Valenzuela-Fernández et al., 2008). In contrast, acetyl transferase α-Tat1 (or Mec-17; Akella et al., 2010; Shida et al., 2010) specifically acetylates α-tubulin K40 (Fig. 3), thus providing a more specific tool to investigate the functions of K40 acetylation. Knockout mice of α-Tat1 are completely void of K40-acetylated tubulin; however, they show only slight phenotypic aberrations, for instance, in their sperm flagellum (Kalebic et al., 2013). A more detailed analysis of α-Tat1 knockout mice demonstrated that absence of K40 acetylation leads to reduced contact inhibition in proliferating cells (Aguilar et al., 2014). In migrating cells, α-Tat1 is targeted to microtubules at the leading edge by clathrin-coated pits, resulting in locally restricted acetylation of those microtubules (Montagnac et al., 2013). A recent structural study of α-Tat1 demonstrated that the low catalytic rate of this enzyme, together with its localization inside the microtubules, caused acetylation to accumulate selectively in stable, long-lived microtubules (Szyk et al., 2014), thus explaining the link between this PTM and stable microtubules in cells. However, the direct cellular function of K40 acetylation on microtubules is still unclear.Open in a separate windowFigure 3.Enzymes involved in PTM of tubulin. Schematic representation of known enzymes (mammalian enzymes are shown) involved in the generation and removal of PTMs shown in Fig. 1. Note that some enzymes still remain unknown, and some modifications are irreversible. (*CCP5 preferentially removes branching points [Rogowski et al., 2010]; however, the enzyme can also hydrolyze linear glutamate chains [Berezniuk et al., 2013]).Recent discoveries have brought up the possibility that tubulin could be subject to multiple acetylation events. A whole-acetylome study identified >10 novel sites on α- and β-tubulin (Choudhary et al., 2009); however, none of these sites have been confirmed. Another acetylation event has been described at lysine 252 (K252) of β-tubulin. This modification is catalyzed by the acetyltransferase San (Fig. 3) and might regulate the assembly efficiency of microtubules as a result of its localization at the polymerization interface (Chu et al., 2011).

Tubulin detyrosination.

Most α-tubulin genes in different species encode a C-terminal tyrosine residue (Fig. 2; Valenzuela et al., 1981). This tyrosine can be enzymatically removed (Hallak et al., 1977) and religated (Fig. 3; Arce et al., 1975). Mapping of tyrosinated and detyrosinated microtubules in cells using specific antibodies (Gundersen et al., 1984; Geuens et al., 1986; Cambray-Deakin and Burgoyne, 1987a) revealed that subsets of interphase and mitotic spindle microtubules are detyrosinated (Gundersen and Bulinski, 1986). As detyrosination was mostly found on stable and long-lived microtubules, especially in neurons (Cambray-Deakin and Burgoyne, 1987b; Robson and Burgoyne, 1989; Brown et al., 1993), it was assumed that this modification promotes microtubule stability (Gundersen et al., 1987; Sherwin et al., 1987). Although a direct stabilization of the microtubule lattice was considered unlikely (Khawaja et al., 1988), it was found more recently that detyrosination protects cellular microtubules from the depolymerizing activity of kinesin-13–type motor proteins, such as KIF2 or MCAK, thus increasing their longevity (Peris et al., 2009; Sirajuddin et al., 2014).Besides kinesin-13 motors, plus end–tracking proteins with cytoskeleton-associated protein glycine-rich (CAP-Gly) domains, such as CLIP170 or p150/glued, specifically interact with tyrosinated microtubules (Peris et al., 2006; Bieling et al., 2008) via this domain (Honnappa et al., 2006). In contrast, kinesin-1 moves preferentially on detyrosinated microtubules tracks in cells (Liao and Gundersen, 1998; Kreitzer et al., 1999; Konishi and Setou, 2009). The effect of detyrosination on kinesin-1 motor behavior was recently measured in vitro, and a small but significant increase in the landing rate and processivity of the motor has been found (Kaul et al., 2014). Such subtle changes in the motor behavior could, in conjunction with other factors, such as regulatory MAPs associated with cargo transport complexes (Barlan et al., 2013), lead to a preferential use of detyrosinated microtubules by kinesin-1 in cells.Despite the early biochemical characterization of a detyrosinating activity, the carboxypeptidase catalyzing detyrosination of α-tubulin has yet to be identified (Hallak et al., 1977; Argaraña et al., 1978, 1980). In contrast, the reverse enzyme, tubulin tyrosine ligase (TTL; Fig. 3; Raybin and Flavin, 1975; Deanin and Gordon, 1976; Argaraña et al., 1980), has been purified (Schröder et al., 1985) and cloned (Ersfeld et al., 1993). TTL modifies nonpolymerized tubulin dimers exclusively. This selectivity is determined by the binding interface between the TTL and tubulin dimers (Szyk et al., 2011, 2013; Prota et al., 2013). In contrast, the so far unidentified detyrosinase acts preferentially on polymerized microtubules (Kumar and Flavin, 1981; Arce and Barra, 1983), thus modifying a select population of microtubules within cells (Gundersen et al., 1987).In most organisms, only one unique gene for TTL exists. Consequently, TTL knockout mice show a huge accumulation of detyrosinated and particularly Δ2-tubulin (see next section). TTL knockout mice die before birth (Erck et al., 2005) with major developmental defects in the nervous system that might be related to aberrant neuronal differentiation (Marcos et al., 2009). TTL is strictly tubulin specific (Prota et al., 2013), indicating that all observed defects in TTL knockout mice are directly related to the deregulation of the microtubule cytoskeleton.

Δ2-tubulin and further C-terminal modification.

A biochemical study of brain tubulin revealed that ∼35% of α-tubulin cannot be retyrosinated (Paturle et al., 1989) because of the lack of the penultimate C-terminal glutamate residue of the primary protein sequence (Fig. 2; Paturle-Lafanechère et al., 1991). This so-called Δ2-tubulin (for two C-terminal amino acids missing) cannot undergo retyrosination as a result of structural constraints within TTL (Prota et al., 2013) and thus is considered an irreversible PTM.Δ2-tubulin accumulates in long-lived microtubules of differentiated neurons, axonemes of cilia and flagella, and also in cellular microtubules that have been artificially stabilized, for instance, with taxol (Paturle-Lafanechère et al., 1994). The generation of Δ2-tubulin requires previous detyrosination of α-tubulin; thus, the levels of this PTM are indirectly regulated by the detyrosination/retyrosination cycle. This mechanistic link is particularly apparent in the TTL knockout mice, which show massive accumulation of Δ2-tubulin in all tested tissues (Erck et al., 2005). Loss of TTL and the subsequent increase of Δ2-tubulin levels were also linked to tumor growth and might contribute to the aggressiveness of the tumors by an as-yet-unknown mechanism (Lafanechère et al., 1998; Mialhe et al., 2001). To date, no specific biochemical role of Δ2-tubulin has been determined; thus, one possibility is that the modification simply locks tubulin in the detyrosinated state.The enzymes responsible for Δ2-tubulin generation are members of a family of cytosolic carboxypeptidases (CCPs; Fig. 3; Kalinina et al., 2007; Rodriguez de la Vega et al., 2007), and most of them also remove polyglutamylation from tubulin (see next section; Rogowski et al., 2010). These enzymes are also able to generate Δ3-tubulin (Fig. 1 A; Berezniuk et al., 2012), indicating that further degradation of the tubulin C-terminal tails are possible; however, the functional significance of this event is unknown.

Polyglutamylation.

Polyglutamylation is a PTM that occurs when secondary glutamate side chains are formed on γ-carboxyl groups of glutamate residues in a protein (Fig. 1, A and B). The modification was first discovered on α- and β-tubulin from the brain (Eddé et al., 1990; Alexander et al., 1991; Rüdiger et al., 1992; Mary et al., 1994) as well as on axonemal tubulin from different species (Mary et al., 1996, 1997); however, it is not restricted to tubulin (Regnard et al., 2000; van Dijk et al., 2008). Using a glutamylation-specific antibody, GT335 (Wolff et al., 1992), it was observed that tubulin glutamylation increases during neuronal differentiation (Audebert et al., 1993, 1994) and that axonemes of cilia and flagella (Fouquet et al., 1994), as well as centrioles of mammalian centrosomes (Bobinnec et al., 1998), are extensively glutamylated.Enzymes catalyzing polyglutamylation belong to the TTL-like (TTLL) family (Regnard et al., 2003; Janke et al., 2005). In mammals, nine glutamylases exist, each of them showing intrinsic preferences for modifying either α- or β-tubulin as well as for initiating or elongating glutamate chains (Fig. 3; van Dijk et al., 2007). Two of the six well-characterized TTLL glutamylases also modify nontubulin substrates (van Dijk et al., 2008).Knockout or depletion of glutamylating enzymes in different model organisms revealed an evolutionarily conserved role of glutamylation in cilia and flagella. In motile cilia, glutamylation regulates beating behavior (Janke et al., 2005; Pathak et al., 2007; Ikegami et al., 2010) via the regulation of flagellar dynein motors (Kubo et al., 2010; Suryavanshi et al., 2010). Despite the expression of multiple glutamylases in ciliated cells and tissues, depletion or knockout of single enzymes often lead to ciliary defects, particularly in motile cilia (Ikegami et al., 2010; Vogel et al., 2010; Bosch Grau et al., 2013; Lee et al., 2013), suggesting essential and nonredundant regulatory functions of these enzymes in cilia.Despite the enrichment of polyglutamylation in neuronal microtubules (Audebert et al., 1993, 1994), knockout of TTLL1, the major polyglutamylase in brain (Janke et al., 2005), did not show obvious neuronal defects in mice (Ikegami et al., 2010; Vogel et al., 2010). This suggests a tolerance of neuronal microtubules to variations in polyglutamylation.Deglutamylases, the enzymes that reverse polyglutamylation, were identified within a novel family of CCPs (Kimura et al., 2010; Rogowski et al., 2010). So far, three out of six mammalian CCPs have been shown to cleave C-terminal glutamate residues, thus catalyzing both the reversal of polyglutamylation and the removal of gene-encoded glutamates from the C termini of proteins (Fig. 3). The hydrolysis of gene-encoded glutamate residues is not restricted to tubulin, in which it generates Δ2- and Δ3-tubulin, but has also been reported for other proteins such as myosin light chain kinase (Rusconi et al., 1997; Rogowski et al., 2010). One enzyme of the CCP family, CCP5, preferentially removes branching points generated by glutamylation, thus allowing the complete reversal of the polyglutamylation modification (Kimura et al., 2010; Rogowski et al., 2010). However, CCP5 can also hydrolyze C-terminal glutamate residues from linear peptide chains similar to other members of the CCP family (Berezniuk et al., 2013).CCP1 is mutated in a well-established mouse model for neurodegeneration, the pcd (Purkinje cell degeneration) mouse (Mullen et al., 1976; Greer and Shepherd, 1982; Fernandez-Gonzalez et al., 2002). The absence of a key deglutamylase leads to strong hyperglutamylation in brain regions that undergo degeneration, such as the cerebellum and the olfactory bulb (Rogowski et al., 2010). When glutamylation levels were rebalanced by depletion or knockout of the major brain polyglutamylase TTLL1 (Rogowski et al., 2010; Berezniuk et al., 2012), Purkinje cells survived. Although the molecular mechanisms of hyperglutamylation-induced degeneration remain to be elucidated, perturbation of neuronal transport, as well as changes in the dynamics and stability of microtubules, is expected to be induced by hyperglutamylation. Increased polyglutamylation levels have been shown to affect kinesin-1–mediated transport in cultured neurons (Maas et al., 2009), and the turnover of microtubules can also be regulated by polyglutamylation via the activation of microtubule-severing enzymes such as spastin (Lacroix et al., 2010).Subtle differences in polyglutamylation can be seen on diverse microtubules in different cell types. The functions of these modifications remain to be studied; however, its wide distribution strengthens the idea that it could be involved in fine-tuning a range of microtubule functions.

Polyglycylation.

Tubulin polyglycylation or glycylation, like polyglutamylation, generates side chains of glycine residues within the C-terminal tails of α- and β-tubulin (Fig. 1, A and B). The modification sites of glycylation are considered to be principally the same as for glutamylation, and indeed, both PTMs have been shown to be interdependent in cells (Rogowski et al., 2009; Wloga et al., 2009). Initially discovered on Paramecium tetraurelia tubulin (Redeker et al., 1994), glycylation has been extensively studied using two antibodies, TAP952 and AXO49 (Bressac et al., 1995; Levilliers et al., 1995; Bré et al., 1996). In contrast to polyglutamylation, glycylation is restricted to cilia and flagella in most organisms analyzed so far.Glycylating enzymes are also members of the TTLL family, and homologues of these enzymes have so far been found in all organisms with proven glycylation of ciliary axonemes (Rogowski et al., 2009; Wloga et al., 2009). In mammals, initiating (TTLL3 and TTLL8) and elongating (TTLL10) glycylases work together to generate polyglycylation (Fig. 3). In contrast, the two TTLL3 orthologues from Drosophila melanogaster can both initiate and elongate glycine side chains (Rogowski et al., 2009).In mice, motile ependymal cilia in brain ventricles acquire monoglycylation upon maturation, whereas polyglycylation is observed only after several weeks (Bosch Grau et al., 2013). Sperm flagella, in contrast, acquire long glycine chains much faster, suggesting that the extent of polyglycylation could correlate with the length of the axonemes (Rogowski et al., 2009). Depletion of glycylases in mice (ependymal cilia; Bosch Grau et al., 2013), zebrafish (Wloga et al., 2009; Pathak et al., 2011), Tetrahymena thermophila (Wloga et al., 2009), and D. melanogaster (Rogowski et al., 2009) consistently led to ciliary disassembly or severe ciliary defects. How glycylation regulates microtubule functions remains unknown; however, the observation that glycylation-depleted axonemes disassemble after initial assembly (Rogowski et al., 2009; Bosch Grau et al., 2013) suggests a role of this PTM in stabilizing axonemal microtubules. Strikingly, human TTLL10 is enzymatically inactive; thus, humans have lost the ability to elongate glycine side chains (Rogowski et al., 2009). This suggests that the elongation of the glycine side chains is not an essential aspect of the function of this otherwise critical tubulin PTM.

Other tubulin PTMs.

Several other PTMs have been found on tubulin. Early studies identified tubulin phosphorylation (Eipper, 1974; Gard and Kirschner, 1985; Díaz-Nido et al., 1990); however, no specific functions were found. The perhaps best-studied phosphorylation event on tubulin takes place at serine S172 of β-tubulin (Fig. 1 A), is catalyzed by the Cdk1 (Fig. 3), and might regulate microtubule dynamics during cell division (Fourest-Lieuvin et al., 2006; Caudron et al., 2010). Tubulin can be also modified by the spleen tyrosine kinase Syk (Fig. 3; Peters et al., 1996), which might play a role in immune cells (Faruki et al., 2000; Sulimenko et al., 2006) and cell division (Zyss et al., 2005; Sulimenko et al., 2006).Polyamination has recently been discovered on brain tubulin (Song et al., 2013), after having been overlooked for many years as a result of the low solubility of polyaminated tubulin. Among several glutamine residues of α- and β-tubulin that can be polyaminated, Q15 of β-tubulin is considered the primary modification site (Fig. 1 A). Polyamination is catalyzed by transglutaminases (Fig. 3), which modify free tubulin as well as microtubules in an irreversible manner, and most likely contribute to the stabilization of microtubules (Song et al., 2013).Tubulin was also reported to be palmitoylated (Caron, 1997; Ozols and Caron, 1997; Caron et al., 2001), ubiquitinated (Ren et al., 2003; Huang et al., 2009; Xu et al., 2010), glycosylated (Walgren et al., 2003; Ji et al., 2011), arginylated (Wong et al., 2007), methylated (Xiao et al., 2010), and sumoylated (Rosas-Acosta et al., 2005). These PTMs have mostly been reported without follow-up studies, and some of them are only found in specific cell types or organisms and/or under specific metabolic conditions. Further studies will be necessary to gain insights into their potential roles for the regulation of the microtubule cytoskeleton.

Current advances and future perspectives

The molecular heterogeneity of microtubules, generated by the expression of different tubulin isotypes and by the PTM of tubulin has fascinated the scientific community for ∼40 years. Although many important advances have been made in the past decade, the dissection of the molecular mechanisms and a comprehensive understanding of the biological functions of tubulin isotypes and PTMs will be a challenging field of research in the near future.

Direct measurements of the impact of tubulin heterogeneity.

The most direct and reliable type of experiments to determine the impact of tubulin heterogeneity on microtubule behavior are in vitro measurements with purified proteins. However, most biophysical work on microtubules has been performed with tubulin purified from bovine, ovine, or porcine brains, which can be obtained in large quantities and with a high degree of purity and activity (Vallee, 1986; Castoldi and Popov, 2003). Brain tubulin is a mixture of different tubulin isotypes and is heavily posttranslationally modified and thus inept for investigating the functions of tubulin heterogeneity (Denoulet et al., 1986; Cambray-Deakin and Burgoyne, 1987b; Paturle et al., 1989; Eddé et al., 1990). Thus, pure, recombinant tubulin will be essential to dissect the roles of different tubulin isoforms and PTMs.Attempts to produce recombinant, functional α- and β-tubulin in bacteria have failed so far (Yaffe et al., 1988), most likely because of the absence of the extensive tubulin-specific folding machinery (Yaffe et al., 1992; Gao et al., 1993; Tian et al., 1996; Vainberg et al., 1998) in prokaryotes. An alternative source of tubulin with less isotype heterogeneity and with almost no PTMs is endogenous tubulin from cell lines such as HeLa, which in the past has been purified using a range of biochemical procedures (Bulinski and Borisy, 1979; Weatherbee et al., 1980; Farrell, 1982; Newton et al., 2002; Fourest-Lieuvin, 2006). Such tubulin can be further modified with tubulin-modifying enzymes, such as polyglutamylases, either by expressing those enzymes in the cells before tubulin purification (Lacroix and Janke, 2011) or in vitro with purified enzymes (Vemu et al., 2014). Despite some technical limitations of these methods, HeLa tubulin modified in cells has been successfully used in an in vitro study on the role of polyglutamylation in microtubule severing (Lacroix et al., 2010).Naturally occurring variants of tubulin isotypes and PTMs can be purified from different organisms, organs, or cell types, but obviously, only some combinations of tubulin isotypes and PTMs can be obtained by this approach. The recent development of an affinity purification method using the microtubule-binding TOG (tumor overexpressed gene) domain of yeast Stu2p has brought a new twist to this approach, as it allows purifying small amounts of tubulin from any cell type or tissue (Widlund et al., 2012).The absence of tubulin heterogeneity in yeast has made budding and fission yeast potential expression systems for recombinant, PTM-free tubulin (Katsuki et al., 2009; Drummond et al., 2011; Johnson et al., 2011). However, the expression of mammalian tubulin in this system has remained impossible. This problem was then partially circumvented by expressing tubulin chimeras that consist of a yeast tubulin body fused to mammalian C-terminal tubulin tails, thus mimicking different tubulin isotypes (Sirajuddin et al., 2014). Moreover, detyrosination can be generated by deleting the key C-terminal residue from endogenous or chimeric α-tubulin (Badin-Larçon et al., 2004), and polyglutamylation is generated by chemically coupling glutamate side chains to specifically engineered tubulin chimeras (Sirajuddin et al., 2014). These approaches allowed the first direct measurements of the impact of tubulin isotypes and PTMs on the behavior of molecular motors in vitro (Sirajuddin et al., 2014) and the analysis of the effects of tubulin heterogeneity on microtubule behavior and interactions inside the yeast cell (Badin-Larçon et al., 2004; Aiken et al., 2014).Currently, the most promising development has been the successful purification of fully functional recombinant tubulin from the baculovirus expression system (Minoura et al., 2013). Using this system, defined α/β-tubulin dimers can be obtained using two different epitope tags on α- and β-tubulin, respectively. Although these epitope tags are essential for separating recombinant from the endogenous tubulin, they could also affect tubulin assembly or microtubule–MAP interactions. Thus, future developments should focus on eliminating these tags.Current efforts have brought the possibility of producing recombinant tubulin into reach. Further improvement and standardization of these methods will certainly provide a breakthrough in understanding the mechanisms by which tubulin heterogeneity contributes to microtubule functions.

Complexity of tubulin—understanding the regulatory principles.

The diversity of tubulin genes (isotypes) and the complexity of tubulin PTMs have led to the proposal of the term “tubulin code” (Verhey and Gaertig, 2007; Wehenkel and Janke, 2014), in analogy to the previously coined histone code (Jenuwein and Allis, 2001). Tubulin molecules consist of a highly structured and thus evolutionarily conserved tubulin body and the unstructured and less conserved C-terminal tails (Nogales et al., 1998). As PTMs and sequence variations within the tubulin body are expected to affect the conserved tubulin fold and therefore the properties of the microtubule lattice, they are not likely to be involved in generating the tubulin code. In contrast, modulations of the C-terminal tails could encode signals on the microtubule surface without perturbing basic microtubule functions and properties (Figs. 1 A and and4).4). Indeed, the highest degree of gene-encoded diversity (Fig. 2) and the highest density and complexity of PTMs (Fig. 1) are found within these tail domains.Open in a separate windowFigure 4.Molecular components of the tubulin code. Schematic representation of potential coding elements that could generate specific signals for the tubulin code. (A) The length of the C-terminal tails of different tubulin isotypes differ significantly (Fig. 2) and could have an impact on the interactions between microtubules and MAPs. (B) Tubulin C-terminal tails are rich in charged amino acid residues. The distribution of these residues and local densities of charges could influence the electrostatic interactions with the tails and the readers. (C) Although each glutamate residue within the C-terminal tails could be considered a potential modification site, only some sites have been found highly occupied in tubulin purifications from native sources. This indicates selectivity of the modification reactions, which can participate in the generation of specific modification patterns (see D). Modification sites might be distinguished by their neighboring amino acid residues, which could create specific modification epitopes. (D) As a result of the large number of modification sites and the variability of side chains, a large variety of modification patterns could be generated within a single C-terminal tail of tubulin. (E) Modification patterns as shown in D can be distinct between α- and β-tubulin. These modification patterns could be differentially distributed at the surface of the microtubule lattice, thus generating a higher-order patterning. Tub, tubulin. For color coding, see Fig. 2.Considering the number of tubulin isotypes plus all potential combinations of PTMs (e.g., each glutamate residue within the C-terminal tubulin tail could be modified by either polyglutamylation or polyglycylation, each of them generating side chains of different lengths; Fig. 4), the number of distinct signals generated by the potential tubulin code would be huge. However, as many of these potential signals represent chemical structures that are similar and might not be reliably distinguished by readout mechanisms, it is possible that the tubulin code generates probabilistic signals. In this scenario, biochemically similar modifications would have similar functional readouts, and marginal differences between those signals would only bias biological processes but not determine them. This stands in contrast to the concept of the histone code, in which precise patterns of different PTMs on the histone proteins encode distinct biological signals.The concept of probabilistic signaling is already inscribed in the machinery that generates the tubulin code. Polyglutamylases and polyglycylases from the TTLL family have preferential activities for either α- or β-tubulin and for generating different lengths of the branched glutamate or glycine chains. Although under conditions of low enzyme concentrations, as found in most cells and tissues, the enzymes seem to selectively generate their preferential type of PTM, higher enzyme concentrations induce a more promiscuous behavior, leading, for instance, to a loss of selectivity for α- or β-tubulin (van Dijk et al., 2007). Similarly, the modifying enzymes might prefer certain modification sites within the C-terminal tails of tubulin but might be equally able to modify other sites, which could be locally regulated in cells. For example, β-tubulin isotypes isolated from mammalian brain were initially found to be glutamylated on single residues (Alexander et al., 1991; Rüdiger et al., 1992), which in the light of the comparably low sensitivity of mass spectrometry at the time might rather indicate a preferential than a unique modification of these sites. Nevertheless, the neuron-specific polyglutamylase for β-tubulin TTLL7 (Ikegami et al., 2006) can incorporate glutamate onto many more modification sites of β-tubulin in vitro (Mukai et al., 2009), which clearly indicates that not all of the possible modification events take place under physiological conditions.Several examples supporting a probabilistic signaling mode of the tubulin code are found in the recent literature. In T. thermophila, a ciliate without tubulin isotype diversity (Gaertig et al., 1993) but with a huge repertoire of tubulin PTMs and tubulin-modifying enzymes (Janke et al., 2005), tubulin can be easily mutagenized to experimentally eliminate sites for PTMs. Mutagenesis of the most commonly occupied glutamylation/glycylation sites within the β-tubulin tails did not generate a clear decrease of glycylation levels nor did it cause obvious phenotypic alterations. This indicates that the modifying enzymes can deviate toward alternative modification sites and that similar PTMs on different sites can compensate the functions of the mutated site. However, when all of the key modification sites were mutated, glycylation became prominently decreased, which led to severe phenotypes, including lethality (Xia et al., 2000). Most strikingly, these phenotypes could be recovered by replacing the C-terminal tail of α-tubulin with the nonmutated β-tubulin tail. This α–β-tubulin chimera became overglycylated and functionally compensated for the absence of modification sites on β-tubulin. The conclusion of this study is that PTM- and isotype-generated signals can fulfill a biological function within a certain range of tolerance.But how efficient is such compensation? The answer can be found in a variety of already described deletion mutants for tubulin-modifying enzymes in different model organisms. Most single-gene knockouts for TTLL genes (glutamylases or glycylases) did not result in prominent phenotypic alterations in mice, even for enzymes that are ubiquitously expressed. Only some highly specialized microtubule structures show functional aberrations upon the deletion of a single enzyme. These “tips of the iceberg” are usually the motile cilia and sperm flagella, which carry very high levels of polyglutamylation and polyglycylation (Bré et al., 1996; Kann et al., 1998; Rogowski et al., 2009). It thus appears that some microtubules are essentially dependent on the generation of specific PTM patterns, whereas others can tolerate changes and appear to function normally. How “normal” these functions are remains to be investigated in future studies. It is possible that defects are subtle and thus overlooked but could become functionally important under specific conditions.A tubulin code also requires readout mechanisms. The most likely “readers” of the tubulin code are MAPs and molecular motors. Considering the probabilistic signaling hypothesis, the expected effects of the signals would be in most cases rather gradual changes, for instance, to fine-tune molecular motor traffic and/or to bias motors toward defined microtubule tracks but not to obliterate motor activity or MAP binding to microtubules. An in vitro study using recombinant tubulin chimeras purified from yeast confirmed this notion (Sirajuddin et al., 2014). By analyzing which elements of the tubulin code can regulate the velocity and processivity of the molecular motors kinesin and dynein, these researchers found that the C-terminal tails of α- and β-tubulin differentially influence the kinetic parameters of the tested motors; however, the modulation was rather modest. One of their striking observations was that a single lysine residue, present in the C-terminal tails of two β-tubulin isotypes (Figs. 2 and and4),4), significantly affected motor traffic and that this effect can be counterbalanced by polyglutamylation. These observations are the first in vitro evidence for the interdependence of different elements of the tubulin code and provide another indication for its probabilistic mode of signaling.

Future directions.

One of the greatest technological challenges to understanding the function of the tubulin code is to detect and interpret subtle and complex regulatory events generated by this code. It will thus be instrumental to further develop tools to better distinguish graded changes in PTM levels on microtubules in cells and tissues (Magiera and Janke, 2013) and to reliably measure subtle modulations of microtubule behavior in reconstituted systems.The current advances in the field and especially the availability of whole-organism models, as well as first insights into the pathological role of tubulin mutations (Tischfield et al., 2011), are about to transform our way of thinking about the regulation of microtubule cytoskeleton. Tubulin heterogeneity generates complex probabilistic signals that cannot be clearly attributed to single biological functions in most cases and that are not essential for most cellular processes. Nevertheless, it has been conserved throughout evolution of eukaryotes and can hardly be dismissed as not important. To understand the functional implications of these processes, we might be forced to reconsider how we define biologically important events and how we measure events that might encode probabilistic signals. The answers to these questions could provide novel insights into how complex systems, such as cells and organisms, are sustained throughout difficult and challenging life cycles, resist to environmental stress and diseases, and have the flexibility needed to succeed in evolution.  相似文献   

10.
11.
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   

12.
13.
14.
To investigate sepal/petal/lip formation in Oncidium Gower Ramsey, three paleoAPETALA3 genes, O. Gower Ramsey MADS box gene5 (OMADS5; clade 1), OMADS3 (clade 2), and OMADS9 (clade 3), and one PISTILLATA gene, OMADS8, were characterized. The OMADS8 and OMADS3 mRNAs were expressed in all four floral organs as well as in vegetative leaves. The OMADS9 mRNA was only strongly detected in petals and lips. The mRNA for OMADS5 was only strongly detected in sepals and petals and was significantly down-regulated in lip-like petals and lip-like sepals of peloric mutant flowers. This result revealed a possible negative role for OMADS5 in regulating lip formation. Yeast two-hybrid analysis indicated that OMADS5 formed homodimers and heterodimers with OMADS3 and OMADS9. OMADS8 only formed heterodimers with OMADS3, whereas OMADS3 and OMADS9 formed homodimers and heterodimers with each other. We proposed that sepal/petal/lip formation needs the presence of OMADS3/8 and/or OMADS9. The determination of the final organ identity for the sepal/petal/lip likely depended on the presence or absence of OMADS5. The presence of OMADS5 caused short sepal/petal formation. When OMADS5 was absent, cells could proliferate, resulting in the possible formation of large lips and the conversion of the sepal/petal into lips in peloric mutants. Further analysis indicated that only ectopic expression of OMADS8 but not OMADS5/9 caused the conversion of the sepal into an expanded petal-like structure in transgenic Arabidopsis (Arabidopsis thaliana) plants.The ABCDE model predicts the formation of any flower organ by the interaction of five classes of homeotic genes in plants (Yanofsky et al., 1990; Jack et al., 1992; Mandel et al., 1992; Goto and Meyerowitz, 1994; Jofuku et al., 1994; Pelaz et al., 2000, 2001; Theißen and Saedler, 2001; Pinyopich et al., 2003; Ditta et al., 2004; Jack, 2004). The A class genes control sepal formation. The A, B, and E class genes work together to regulate petal formation. The B, C, and E class genes control stamen formation. The C and E class genes work to regulate carpel formation, whereas the D class gene is involved in ovule development. MADS box genes seem to have a central role in flower development, because most ABCDE genes encode MADS box proteins (Coen and Meyerowitz, 1991; Weigel and Meyerowitz, 1994; Purugganan et al., 1995; Rounsley et al., 1995; Theißen and Saedler, 1995; Theißen et al., 2000; Theißen, 2001).The function of B group genes, such as APETALA3 (AP3) and PISTILLATA (PI), has been thought to have a major role in specifying petal and stamen development (Jack et al., 1992; Goto and Meyerowitz, 1994; Krizek and Meyerowitz, 1996; Kramer et al., 1998; Hernandez-Hernandez et al., 2007; Kanno et al., 2007; Whipple et al., 2007; Irish, 2009). In Arabidopsis (Arabidopsis thaliana), mutation in AP3 or PI caused identical phenotypes of second whorl petal conversion into a sepal structure and third flower whorl stamen into a carpel structure (Bowman et al., 1989; Jack et al., 1992; Goto and Meyerowitz, 1994). Similar homeotic conversions for petal and stamen were observed in the mutants of the AP3 and PI orthologs from a number of core eudicots such as Antirrhinum majus, Petunia hybrida, Gerbera hybrida, Solanum lycopersicum, and Nicotiana benthamiana (Sommer et al., 1990; Tröbner et al., 1992; Angenent et al., 1993; van der Krol et al., 1993; Yu et al., 1999; Liu et al., 2004; Vandenbussche et al., 2004; de Martino et al., 2006), from basal eudicot species such as Papaver somniferum and Aquilegia vulgaris (Drea et al., 2007; Kramer et al., 2007), as well as from monocot species such as Zea mays and Oryza sativa (Ambrose et al., 2000; Nagasawa et al., 2003; Prasad and Vijayraghavan, 2003; Yadav et al., 2007; Yao et al., 2008). This indicated that the function of the B class genes AP3 and PI is highly conserved during evolution.It has been thought that B group genes may have arisen from an ancestral gene through multiple gene duplication events (Doyle, 1994; Theißen et al., 1996, 2000; Purugganan, 1997; Kramer et al., 1998; Kramer and Irish, 1999; Lamb and Irish, 2003; Kim et al., 2004; Stellari et al., 2004; Zahn et al., 2005; Hernandez-Hernandez et al., 2007). In the gymnosperms, there was a single putative B class lineage that duplicated to generate the paleoAP3 and PI lineages in angiosperms (Kramer et al., 1998; Theißen et al., 2000; Irish, 2009). The paleoAP3 lineage is composed of AP3 orthologs identified in lower eudicots, magnolid dicots, and monocots (Kramer et al., 1998). Genes in this lineage contain the conserved paleoAP3- and PI-derived motifs in the C-terminal end of the proteins, which have been thought to be characteristics of the B class ancestral gene (Kramer et al., 1998; Tzeng and Yang, 2001; Hsu and Yang, 2002). The PI lineage is composed of PI orthologs that contain a highly conserved PI motif identified in most plant species (Kramer et al., 1998). Subsequently, there was a second duplication at the base of the core eudicots that produced the euAP3 and TM6 lineages, which have been subject to substantial sequence changes in eudicots during evolution (Kramer et al., 1998; Kramer and Irish, 1999). The paleoAP3 motif in the C-terminal end of the proteins was retained in the TM6 lineage and replaced by a conserved euAP3 motif in the euAP3 lineage of most eudicot species (Kramer et al., 1998). In addition, many lineage-specific duplications for paleoAP3 lineage have occurred in plants such as orchids (Hsu and Yang, 2002; Tsai et al., 2004; Kim et al., 2007; Mondragón-Palomino and Theißen, 2008, 2009; Mondragón-Palomino et al., 2009), Ranunculaceae, and Ranunculales (Kramer et al., 2003; Di Stilio et al., 2005; Shan et al., 2006; Kramer, 2009).Unlike the A or C class MADS box proteins, which form homodimers that regulate flower development, the ability of B class proteins to form homodimers has only been reported in gymnosperms and in the paleoAP3 and PI lineages of some monocots. For example, LMADS1 of the lily Lilium longiflorum (Tzeng and Yang, 2001), OMADS3 of the orchid Oncidium Gower Ramsey (Hsu and Yang, 2002), and PeMADS4 of the orchid Phalaenopsis equestris (Tsai et al., 2004) in the paleoAP3 lineage, LRGLOA and LRGLOB of the lily Lilium regale (Winter et al., 2002), TGGLO of the tulip Tulipa gesneriana (Kanno et al., 2003), and PeMADS6 of the orchid P. equestris (Tsai et al., 2005) in the PI lineage, and GGM2 of the gymnosperm Gnetum gnemon (Winter et al., 1999) were able to form homodimers that regulate flower development. Proteins in the euAP3 lineage and in most paleoAP3 lineages were not able to form homodimers and had to interact with PI to form heterodimers in order to regulate petal and stamen development in various plant species (Schwarz-Sommer et al., 1992; Tröbner et al., 1992; Riechmann et al., 1996; Moon et al., 1999; Winter et al., 2002; Kanno et al., 2003; Vandenbussche et al., 2004; Yao et al., 2008). In addition to forming dimers, AP3 and PI were able to interact with other MADS box proteins, such as SEPALLATA1 (SEP1), SEP2, and SEP3, to regulate petal and stamen development (Pelaz et al., 2000; Honma and Goto, 2001; Theißen and Saedler, 2001; Castillejo et al., 2005).Orchids are among the most important plants in the flower market around the world, and research on MADS box genes has been reported for several species of orchids during the past few years (Lu et al., 1993, 2007; Yu and Goh, 2000; Hsu and Yang, 2002; Yu et al., 2002; Hsu et al., 2003; Tsai et al., 2004, 2008; Xu et al., 2006; Guo et al., 2007; Kim et al., 2007; Chang et al., 2009). Unlike the flowers in eudicots, the nearly identical shape of the sepals and petals as well as the production of a unique lip in orchid flowers make them a very special plant species for the study of flower development. Four clades (1–4) of genes in the paleoAP3 lineage have been identified in several orchids (Hsu and Yang, 2002; Tsai et al., 2004; Kim et al., 2007; Mondragón-Palomino and Theißen, 2008, 2009; Mondragón-Palomino et al., 2009). Several works have described the possible interactions among these four clades of paleoAP3 genes and one PI gene that are involved in regulating the differentiation and formation of the sepal/petal/lip of orchids (Tsai et al., 2004; Kim et al., 2007; Mondragón-Palomino and Theißen, 2008, 2009). However, the exact mechanism that involves the orchid B class genes remains unclear and needs to be clarified by more experimental investigations.O. Gower Ramsey is a popular orchid with important economic value in cut flower markets. Only a few studies have been reported on the role of MADS box genes in regulating flower formation in this plant species (Hsu and Yang, 2002; Hsu et al., 2003; Chang et al., 2009). An AP3-like MADS gene that regulates both floral formation and initiation in transgenic Arabidopsis has been reported (Hsu and Yang, 2002). In addition, four AP1/AGAMOUS-LIKE9 (AGL9)-like MADS box genes have been characterized that show novel expression patterns and cause different effects on floral transition and formation in Arabidopsis (Hsu et al., 2003; Chang et al., 2009). Compared with other orchids, the production of a large and well-expanded lip and five small identical sepals/petals makes O. Gower Ramsey a special case for the study of the diverse functions of B class MADS box genes during evolution. Therefore, the isolation of more B class MADS box genes and further study of their roles in the regulation of perianth (sepal/petal/lip) formation during O. Gower Ramsey flower development are necessary. In addition to the clade 2 paleoAP3 gene OMADS3, which was previously characterized in our laboratory (Hsu and Yang, 2002), three more B class MADS box genes, OMADS5, OMADS8, and OMADS9, were characterized from O. Gower Ramsey in this study. Based on the different expression patterns and the protein interactions among these four orchid B class genes, we propose that the presence of OMADS3/8 and/or OMADS9 is required for sepal/petal/lip formation. Further sepal and petal formation at least requires the additional presence of OMADS5, whereas large lip formation was seen when OMADS5 expression was absent. Our results provide a new finding and information pertaining to the roles for orchid B class MADS box genes in the regulation of sepal/petal/lip formation.  相似文献   

15.
16.
17.
18.
Proper brain wiring during development is pivotal for adult brain function. Neurons display a high degree of polarization both morphologically and functionally, and this polarization requires the segregation of mRNA, proteins, and lipids into the axonal or somatodendritic domains. Recent discoveries have provided insight into many aspects of the cell biology of axonal development including axon specification during neuronal polarization, axon growth, and terminal axon branching during synaptogenesis.

Introduction

Axon development can be divided into three main steps: (1) axon specification during neuronal polarization, (2) axon growth and guidance, and (3) axon branching and presynaptic differentiation (Fig. 1; Barnes and Polleux, 2009; Donahoo and Richards, 2009). These three steps are exemplified during neocortical development in the mouse: upon neurogenesis, newly born neurons engage long-range migration and polarize (Fig. 1, A and B) by adopting a bipolar morphology with a leading and a trailing process (Fig. 1 C). During migration (approximately from embryonic day [E]11 to E18 in the mouse cortex), the trailing process becomes the axon and extends rapidly while being guided to its final destination (lasts until around postnatal day [P]7 in mouse corticofugal axons with distant targets like the spinal cord; Fig. 1, D–F). Finally, upon reaching its target area, extensive axonal branching occurs during the formation of presynaptic contacts with specific postsynaptic partners (during the second and third postnatal week in the mouse cortex; Fig. 1, G–I). Disruption of any of these steps is thought to lead to various neurodevelopmental disorders ranging from mental retardation and infantile epilepsy to autism spectrum disorders (Zoghbi and Bear, 2012). This review will provide an overview of some of the cellular and molecular mechanisms underlying axon specification, growth, and branching.Open in a separate windowFigure 1.Axon specification, growth, and branching during mouse cortical development. Three stages of the development of callosal axons of cortical pyramidal neurons from the superficial layers 2/3 of the somatosensory cortex in the mouse visualized using long-term in utero cortical electroporation. For this class of model axons, development can be divided in three main stages: (1) neurogenesis and axon specification, occurring mostly at embryonic ages (A–C); (2) axon growth/guidance during the first postnatal week (D–F); and (3) axon branching and synapse formation until approximately the end of the third postnatal week (G–I). A, D, and G show coronal sections of mouse cortex at the indicated ages after in utero cortical electroporation of a GFP-coding plasmid at E15.5 in superficial neuron precursors in one brain hemisphere only (GFP signal in inverted color, dotted line indicates the limits of the brain). B, E, and H are a schematic representation of the main morphological changes observed in callosally projecting axons (red) at the corresponding ages. C shows the typical bipolar morphology of a migrating neuron emitting a trailing process (TP) and a leading process (LP) that will ultimately become the axon and dendrite, respectively. F and I show typical axon projections of layer 2/3 neurons located in the primary somatosensory area at P8 and P21, respectively. Neurons and axons in C, F, and I are visualized by GFP expression (inverted color). Image in C is modified from Barnes et al. (2007) with permission from Elsevier. Images in D, F, G, and I are reprinted from Courchet et al. (2013) with permission from Elsevier.

Neuronal polarization and axon specification

Neuronal polarization is the process of breaking symmetry in the newly born cell to create the asymmetry inherent to the formation of the axonal and somatodendritic compartments (Dotti and Banker, 1987). The mechanisms underlying this process have been studied extensively in vitro and more recently in vivo, but the exact sequence of events has remained elusive (Neukirchen and Bradke, 2011) partly because it is studied in various neuronal cell types that might not use the same extrinsic/intrinsic mechanisms to polarize. It is highly likely that at least three factors underlie neuronal polarization: extracellular cues, intracellular signaling cascades, and subcellular organelle localization. The partition-defective proteins (PARs) are a highly conserved family of proteins including two dyads (Par3/Par6 adaptor proteins and the Par4/Par1 serine/threonine kinases) that are required for polarization and axon formation (Shi et al., 2003, 2004; Barnes et al., 2007; Shelly et al., 2007; Chen et al., 2013), while many other intracellular signaling molecules also support axon formation (Oliva et al., 2006; Rašin et al., 2007; Barnes and Polleux, 2009; Shelly et al., 2010; Cheng et al., 2011; Hand and Polleux, 2011; Cheng and Poo, 2012; Gärtner et al., 2012). These intracellular signaling pathways are influenced by localized extracellular cues that instruct which neurite becomes the axon by either directly promoting axon extension or repressing axon growth in favor of dendritic growth (Adler et al., 2006; Yi et al., 2010; Randlett et al., 2011b; Shelly et al., 2011).The role of organelle subcellular localization during neuronal polarization is a more controversial issue. Initially, the orientation of organelles, including the Golgi complex, centrosomes, mitochondria, and endosomes, was shown to correlate with the neurite that becomes the axon in vitro (Bradke and Dotti, 1997; de Anda et al., 2005, 2010) and in vivo (de Anda et al., 2010). However, more recent studies suggest that the positioning of the centrosome is not necessary for neuronal polarization (Distel et al., 2010; Nguyen et al., 2011). Centrosome localization is likely constrained by microtubule organization within the cell, and therefore the centrosome position within the cell changes dynamically during different stages of polarization (Sakakibara et al., 2013). The question of how the interplay between extracellular cues, intracellular signaling, and organelle localization lead to polarization has pushed the field to perform more extensive in vivo imaging studies as in vitro systems/models have a difficult time recapitulating the complex environment and rely on neurons that were previously polarized in vivo.Like other epithelial cells, neural progenitors present a high degree of polarization along the apico-basal axis (Götz and Huttner, 2005). One of the major questions still needing to be addressed is how, or if, newly born mammalian neurons inherit some level of asymmetry from their parent progenitors (Barnes and Polleux, 2009). Recent studies have attempted to answer this question in vivo but have found that the answer might vary in each neuronal subtype. Retinal ganglion cells (RGCs), retinal bipolar neurons, and tegmental hindbrain nuclei neurons seem to inherit the apical/basolateral polarity from their progenitors (Morgan et al., 2006; Zolessi et al., 2006; Distel et al., 2010; Randlett et al., 2011a). In cortical neurons, hippocampal neurons, and cerebellar granule neurons, this relationship is unclear, in part because newly born cortical neurons first exhibit a multipolar morphology with dynamic neurites emerging from the cell body before adopting a bipolar morphology, suggesting they may not retain a predisposed parental polarity (Hand et al., 2005; Barnes et al., 2007). Other factors also suggest that different neuronal subtypes use different mechanisms during polarization. One such factor is the position where neurons specify their axon relative to the original apical/basolateral axis of their progenitors. As an example, cortical neurons in the mouse brain protrude an axon from the membrane facing the original apical surface toward the ventricular zone (Hand et al., 2005; Barnes et al., 2007; Shelly et al., 2007), whereas zebrafish RGCs form their axon from the membrane on the basolateral side (Zolessi et al., 2006; Randlett et al., 2011b). Another significant difference between cortical neurons and RGCs is related to the timing of axogenesis and dendrogenesis. RGCs tend to form their axons and dendrites at the same time during migration (Zolessi et al., 2006; Randlett et al., 2011b). However, cortical neurons form a long axon during migration before significant dendrite arborization takes place. These differences in the regulation of polarization and sequence of axon versus dendrite outgrowth may be linked to the localization of extracellular cues relative to the immature neuron during polarization (Yi et al., 2010).

Neuronal polarization, cytoskeletal dynamics, and polarized transport

What exactly makes the axonal compartment distinct from the somatodendritic domain? This can most easily be illustrated by focusing on the cytoskeleton that forms the framework of the developing axon. The cytoskeleton is composed of microtubules, actin filaments, and intermediate filaments (also called neurofilaments) along with their associated binding partners. Microtubules are composed of α- and β-tubulin subunits that polymerize to form a long filament intrinsically polarized by the addition of tubulin subunits to only one side of the growing filament called the plus end, while on the opposite side depolymerization occurs. It was discovered more than thirty years ago that the axon of a neuron contains a very uniform distribution of microtubules with the plus end facing away from the cell body (Heidemann et al., 1981). Through the years this observation was confirmed in many neuron cell types, and it was determined that dendrites do not have this uniform plus-end out network of microtubules (Fig. 2; Baas et al., 1988). Dendrites appear to have a complex array of microtubule orientations that may vary between species and/or neuronal subtypes. Current research shows that proximal dendrites are composed of mainly minus-end out microtubules, whereas more distal dendrites transition from an equal distribution of minus-end out and plus-end out microtubules to mainly plus-end out microtubules (Stone et al., 2008; Yin et al., 2011; Ori-McKenney et al., 2012). The orientation of microtubules matters greatly because it determines the relative contribution of microtubule-dependent motor proteins (kinesins and dyneins), which are the main motor proteins carrying various cargoes within cells and in particular are responsible for long-range transport in very large cells such as neurons. Dynein (a minus end–directed microtubule motor) is known to be responsible for both the transport of microtubules away from the cell body and for the uniform polarity of microtubules in the axon (Ahmad et al., 1998; Zheng et al., 2008). Recently, it was discovered that kinesin-1 (a plus end–directed microtubule motor) is required for the minus-end out orientation of microtubules in the dendrites of Caenorhabditis elegans DA9 neurons through selective transport of plus-end out microtubule fragments out of the dendrite (Yan et al., 2013). Another hallmark that differentiates the axonal and somatodendritic compartments is the microtubule-associated proteins (MAPs) that decorate microtubules to regulate their bundling and stability (Hirokawa et al., 2010). Microtubules in the axon are mainly decorated by Tau and MAP1B, whereas microtubules in the dendrites are labeled by proteins of the MAP2a-c family. The role of Tau in axon elongation remains controversial because early reports (Harada et al., 1994; Tint et al., 1998; Dawson et al., 2001) of Tau knockout alone suggested that axons were unaffected, but this apparent lack of phenotype might originate from the functional redundancy between MAPs as Tau/MAP1b double knockout mice show clear axon growth defects (Takei et al., 2000).Open in a separate windowFigure 2.Polarity maintenance and trafficking of somatodendritic and axonal proteins. Neurons are polarized into two main compartments: the somatodendritic domain and the axon. These domains are characterized by the underlying cytoskeleton and their unique protein signatures. The axonal cytoskeleton is defined by its uniform microtubule orientation where each microtubule is oriented with its plus end away from the cell body, while the dendrites contain a mixture of microtubules oriented in both directions. The proximal axon is characterized by a structure known as the axon initial segment (AIS, see inset). This highly ordered structure creates a diffusion barrier between the axonal compartment and the rest of the cell. F-actin is responsible for the cytoplasmic barrier, while sodium channels anchored by Ankyrin G form the basis of the plasma membrane barrier. Tau is retained in the axon by a microtubule-based filter at the AIS. Molecular motors (including kinesin, dynein, and myosin) then use the underlying cytoskeleton to restrict cargo transport to either the axon (such as Cntn1 and L1) or the dendrites (such as PSD95, AMPARs, and NMDARs).The dynamics of actin polymerization into actin filaments (F-actin) also play an important role in defining the axonal compartment, and contain an intrinsic polarity based on the polymerization of the free G-actin subunits (Hirokawa et al., 2010). Beyond the well-documented early role of F-actin dynamics in neurite outgrowth, multiple groups have shown that the disruption of actin polymerization allows dendritically localized proteins to incorrectly enter the axonal compartment (Winckler et al., 1999; Lewis et al., 2009; Song et al., 2009). The existence of a “diffusion barrier” in the proximal part of newly formed axons (Song et al., 2009) was long suspected. One of the current hypotheses is that a dense F-actin meshwork creates a cytoplasmic diffusion barrier shortly after polarization, which in part separates the axonal compartment from the neuronal cell body (Fig. 2, inset). Based on functional analysis and electron microscopy analysis, this “F-actin–based filter” is oriented so that the plus ends point toward the cell body while the minus ends point into the axon (Lewis et al., 2009, 2011; Watanabe et al., 2012). Two recent papers show via high resolution imaging techniques that indeed the axon has a unique F-actin network that is not found in dendrites (Watanabe et al., 2012; Xu et al., 2013). The development of this F-actin meshwork appears to directly precede the formation of the axon initial segment (AIS; Song et al., 2009; Galiano et al., 2012). An intra-axonal diffusion barrier, composed of Spectrins and Ankyrin B, defines the eventual position of the AIS. This boundary excludes Ankyrin G, which instead clusters in the most proximal part of the axon close to the cell body, where the AIS will form (Galiano et al., 2012). Ankyrin G is required for AIS formation and maintenance, and its loss causes the axon to start forming protrusions resembling dendritic spines (Hedstrom et al., 2008). Microtubules also play an important role at the AIS, as recent evidence suggests that Tau is retained in the axon through a microtubule-based diffusion barrier independently of the F-actin based filter (X. Li et al., 2011). The AIS is important in the formation of a plasma membrane barrier between the axonal and somatodendritic domains and its disruption affects both neuronal polarity and function because it is critical for clustering of voltage-dependent sodium channels and action potential initiation (Rasband, 2010).One of the critical cellular mechanisms underlying neuronal polarization is the polarized transport of various cargoes in axons and dendrites. Transport of proteins and various organelles is performed by the microtubule-dependent motor proteins kinesin and dynein (Hirokawa et al., 2010). Studies from many laboratories have demonstrated that kinesin motors can carry cargo to both the axonal and dendritic compartments (Burack et al., 2000; Nakata and Hirokawa, 2003). The mechanism for how selection occurs is not completely understood, but it probably incorporates both the affinity of the kinesin head for microtubules and the cargo bound to the motor protein (Nakata and Hirokawa, 2003; Song et al., 2009; Jenkins et al., 2012). In the axon, dynein works to bring cargo and retrograde signals back to the cell body, whereas in the dendrites it is responsible for much of the transport from the soma into the dendrites (Zheng et al., 2008; Kapitein et al., 2010; Harrington and Ginty, 2013). Additionally, the F-actin–dependent myosin motors can affect the polarized transport of cargos by using the F-actin–based cytoplasmic filter at the AIS to deny or facilitate entry of vesicles into the axon. Loss of the actin filter or myosin Va activity (a plus end–directed motor) allows dendritic cargos into the axon, whereas myosin VI (a minus end–directed motor) both removes axonal proteins from the dendritic surface and funnels vesicles containing axonal proteins through the actin filter at the AIS (Lewis et al., 2009, 2011; Al-Bassam et al., 2012). A current working hypothesis is that vesicles composed of multiple cargoes contain binding sites for each of these motors, and that through unknown mechanisms the activity of the motors can be differentially regulated to control the directionality of transport. An interesting example of how the interplay between different motors and cargo adaptors could lead to polarized transport was recently described for mitochondria (van Spronsen et al., 2013).

Axon growth

Microtubule dynamics regulate axon growth.

After axon specification, axon growth constitutes the second step of axonal development and is tightly linked to axon guidance toward the proper postsynaptic targets. Axon elongation by the growth cone is the product of two opposite forces (Fig. 3): slow axonal transport and the polymerization of microtubules providing a pushing force from the axon shaft, and the retrograde flow of actin providing a pulling force at the front of the growth cone (Letourneau et al., 1987; Suter and Miller, 2011). Although coordinated actin and microtubule dynamics are required for the proper function of the growth cone, it was reported that agents disrupting the actin cytoskeleton have limited consequences on axon elongation and are rather involved in axon guidance in vitro (Marsh and Letourneau, 1984; Ruthel and Hollenbeck, 2000) and in vivo (Bentley and Toroian-Raymond, 1986). Furthermore, local disruption of actin organization in the growth cone of minor neurites allows them to turn into axons (Bradke and Dotti, 1999; Kunda et al., 2001), indicating that the dense actin network present at the periphery of an immature neuronal cell body and in immature neurites may prevent microtubule protrusion and elongation necessary for axon specification.Open in a separate windowFigure 3.Cytoskeletal changes during axon elongation and branching. Representation of axon elongation and collateral branch formation in a cultured neuron. Axon growth is a discontinuous process, and collateral branches often originate from sites where the growth cone paused (gray dotted line), after it has resumed its progression. Other modalities of branch formation can occur through the formation of filopodia and lamellipodia. Red box shows a magnification of the main growth cone. Microtubules from the axon shaft spread into the central (C) zone. Some microtubules pass through the transition (T) zone, containing F-actin arcs, to explore filopodia from the peripheral (P) zone. Upon the proper stimulation by extracellular guidance cues or growth-promoting cues, microtubules are stabilized and invade the P-zone where they provide a pushing force, which, combined with the traction force from the actin treadmilling, provides the force required for growth cone extension. Green box shows the cytoskeletal changes occurring during collateral branch formation in the axon. Filopodia and lamellipodia are primarily F-actin–based protrusions that get invaded by microtubules, then elongate upon microtubule bundling. At later developmental stages, axon branches are stabilized or retracted (blue box) by mechanisms relying on the access to extracellular neurotrophins and/or neuronal activity and synapse formation.Contrary to actin, microtubule polymerization is required to sustain axon elongation and branching (Letourneau et al., 1987; Baas and Ahmad, 1993). Axonal proteins and cytoskeletal elements are transported along the axon through slow axonal transport by molecular motors (Yabe et al., 1999; Xia et al., 2003). It is still controversial whether tubulin and other cytoskeletal elements are transported in the axon as monomers and/or as polymers (Roy et al., 2000; Terada et al., 2000; Wang et al., 2000; Brown, 2003; Terada, 2003). Nonetheless, disruption of the slow transport of tubulin impairs the pushing force resulting from microtubule polymerization and impairs axon elongation (Suter and Miller, 2011). Therefore, it is not surprising that axon growth is affected in vitro and in vivo by disruption of plus-end microtubule-binding proteins such as APC (Shi et al., 2004; Zhou et al., 2004; Yokota et al., 2009; Chen et al., 2011) or EB1 and EB3 (Zhou et al., 2004; Jiménez-Mateos et al., 2005; Geraldo et al., 2008), microtubule-associated proteins such as MAP1B (Black et al., 1994; Takei et al., 2000; Dajas-Bailador et al., 2012; Tortosa et al., 2013), or proteins regulating microtubule severing and reorganization such as KIF2A (Homma et al., 2003), katanin, and spastin (Karabay et al., 2004; Yu et al., 2005; Wood et al., 2006; Butler et al., 2010).The contribution of microtubule dynamics to axon growth is not limited to growth cone dynamics but also involves axonal transport and polymerization along the axon shaft. Moreover, changing the balance between microtubule stabilization and depolymerization by local application of microtubule stabilizing agents is sufficient to instruct one neurite to grow and adopt an axon fate (Witte et al., 2008). Many kinase pathways converge on Tau and other axonal MAPs to regulate their function by phosphorylation (Morris et al., 2011). Among them, the MARK kinases regulate microtubule stability and axonal transport through phosphorylation of Tau (Drewes et al., 1997; Mandelkow et al., 2004). Interestingly, MARK-related kinases SAD-A/B control axon specification in part through phosphorylation of Tau (Barnes et al., 2007) and have very recently been linked to the growth and branching of the axons of sensory neurons (Lilley et al., 2013). Our work recently demonstrated that another family member related to MARKs and SAD kinases, called NUAK1, controls axon branching of mouse cortical neurons through the regulation of presynaptic mitochondria capture (Courchet et al., 2013). To what extent the regulation of Tau and other MAPs by the MARKs, SADs, and NUAK1 kinases contributes to axon elongation remains to be explored.

Where does axon elongation take place?

Growth cone progression and guidance constitute the main driver of axonal growth during development, but this process is unlikely to account for the totality of axon elongation. This is especially true after the axon has reached its final target but the axon shaft keeps growing in proportion to the rest of the body. One mechanism that may contribute to this “interstitial” form of axon elongation during brain/body size increase (see Fig. 1 for an example during postnatal cortex growth) is axon stretching, a process that can induce axon elongation in vitro (Smith et al., 2001; Pfister et al., 2004; Loverde et al., 2011) and in vivo (Abe et al., 2004). Aside from extreme stretching performed through the application of external forces, stretching could also contribute to the natural elongation of the axon in response to the tension resulting from growth cone progression (Suter and Miller, 2011).Axon elongation requires the addition of new lipids, proteins, cytoskeleton elements, and organelles along the axon. Where does the synthesis and incorporation of new components take place? Polysaccharides and cholesterol synthesis mostly occur in the cell body; however, lipid synthesis and/or incorporation can occur along the axon as well (Posse De Chaves et al., 2000; Hayashi et al., 2004). The growth cone is also a site of endocytosis, membrane recycling, and exocytosis (Kamiguchi and Yoshihara, 2001; Winckler and Yap, 2011; Nakazawa et al., 2012). One of the best studied examples of endocytosis and its role in axon growth and neuronal survival is the retrograde trafficking of TrkA receptor by target-derived nerve growth factor (NGF) in the peripheral nervous system (Harrington and Ginty, 2013).

Axon branching and presynaptic differentiation

Where do axon branches form?

The last step of axon development is terminal branching, which allows a single axon to connect to a broad set of postsynaptic targets. Collateral branches are formed along the axon through two distinct mechanisms: the first modality of branching is through splitting or bifurcation of the growth cone, which is linked to axon guidance and to the capacity of one single neuron to reach two targets that are far apart with one single axon. Growth cone splitting is observed in vivo in various neuron types including cortical neurons (Sato et al., 1994; Bastmeyer and O’Leary, 1996; Dent et al., 1999; Tang and Kalil, 2005), sympathetic neurons (Letourneau et al., 1986), motorneurons (Matheson and Levine, 1999), sensory neurons (Ma and Tessier-Lavigne, 2007), or mushroom body neurons in Drosophila (Wang et al., 2002). The second modality, known as interstitial branching, occurs through the formation of collateral branches directly along the axon shaft. Contrary to growth cone splitting, interstitial branching serves the purpose of raising axon coverage locally in order to define their “presynaptic territory”, and may contribute to increased network connectivity (Portera-Cailliau et al., 2005). Although both mechanisms can occur simultaneously in the same neuron, the relative importance of splitting versus interstitial branching is highly divergent from one neuron type to the other (Bastmeyer and O’Leary, 1996; Matheson and Levine, 1999; Portera-Cailliau et al., 2005).In culture, the axon grows in a non-continuous fashion with frequent growth cone pausing. Time-lapse imaging of sensorimotor neurons revealed that interstitial branching often occurs at the site where the growth cone paused, shortly after it has continued its course (Szebenyi et al., 1998). Accordingly, treatments with neurotrophins that slow the growth cone correlate with increased axon branching (Szebenyi et al., 1998). This suggests that growth cone pausing leaves a “mark” along the axon shaft that might predetermine future sites of branching (Kalil et al., 2000). However, local applications of neurotrophins shows that aside from growth cone pause sites the axon shaft remains competent to form collateral branches upon stimulation by extracellular factors (Gallo and Letourneau, 1998; Szebenyi et al., 2001), through the formation of filopodia or lamellipodia. Similar observations in vivo revealed that cortical axons are highly dynamic during development and form multiple filopodia that are the origin of collateral branches (Bastmeyer and O’Leary, 1996). Lamellipodia can be observed as motile, F-actin–dependent “waves” along the axon in vitro (Ruthel and Banker, 1998) and in vivo (Flynn et al., 2009). Moreover, disruption of microtubule organization impairs lamellipodia formation along the axon and is correlated with decreased axon branching (Dent and Kalil, 2001; Tint et al., 2009).

Cytoskeleton dynamics and axon branch formation.

Regardless of what type of protrusion gives rise to a branch, cytoskeletal reorganization in the nascent branch generally follows a similar sequence (Fig. 3): initially F-actin filament reorganization gives rise to a protrusion (filopodia, lamellipodia), followed by microtubule invasion of this otherwise transient structure to consolidate it, before the mature branch starts elongating through microtubule bundling (Gallo, 2011). Actin filaments accumulate along the axon and form “patches” that serve as nucleators for axon protrusions such as filopodia and lamellipodia (Korobova and Svitkina, 2008; Mingorance-Le Meur and O’Connor, 2009; Ketschek and Gallo, 2010). The mRNA for β-actin and regulators of actin polymerization such as Wave1 or Cortactin accumulate along the axons of sensory neurons and form hot-spots of local translation that are associated to NGF-dependent branching (Spillane et al., 2012; Donnelly et al., 2013). Subsequently, microtubules in the axon shaft undergo fragmentation at branch points as a prelude to branch invasion by microtubules (Yu et al., 1994, 2008; Gallo and Letourneau, 1998; Dent et al., 1999; Hu et al., 2012), a process that may disrupt transport locally to help trap molecules and organelles at branch points. Moreover, severed microtubules are transported into branches, a process required for branch stabilization (Gallo and Letourneau, 1999; Ahmad et al., 2006; Qiang et al., 2010; Hu et al., 2012). Interestingly, it is clear that, just like growth cone–mediated axon elongation, F-actin and microtubule reorganization events are interconnected to sustain axon branching (Dent and Kalil, 2001). As an example, microtubule-severing enzymes can also contribute to actin nucleation and filopodia formation (Hu et al., 2012).

Is axon branching linked to axon elongation?

Like in the growth cone, cytoskeleton reorganization constitutes the backbone of branch formation. It is therefore not surprising that many manipulations of the cytoskeleton affect both axon elongation and branch formation (Homma et al., 2003; Chen et al., 2011). Moreover, conditions that primarily disrupt axon elongation could secondarily disrupt branching by impairing the ability of the nascent branch to grow. However, axon elongation and axon branching can be considered as two separate phenomena and can be operationally separated because conditions disrupting one do not systematically affect the other. As an example, the microtubule-severing proteins katanin and spastin have differential consequences on axon elongation (primarily dependent upon katanin function) and branching (mostly spastin mediated; Qiang et al., 2010), taxol treatment (which stabilizes microtubules) affects axon elongation but not branching (Gallo and Letourneau, 1999), and disruption of TrkA endocytosis by knock-down of Unc51-like kinase (ULK1/2) proteins has opposite effects on axon elongation and branching (Zhou et al., 2007). In vivo, superficial layer cortical neurons initially go through a phase of elongation through the corpus callosum without branching (see Fig. 1), then stop elongating and form collateral branches in the contralateral cortex (Mizuno et al., 2007; Wang et al., 2007). It is conceivable that even before myelination, axons are actively prevented from branching at places and stages when they elongate (for example in the white matter of the neocortex) where they tend to be highly fasciculated. The identities of the molecules that inhibit interstitial branching along the axon shaft are currently unknown.

Regulation of axon branching by activity.

Immature neurons display spontaneous activity in the form of calcium waves (Gu et al., 1994; Gomez and Spitzer, 1999; Gomez et al., 2001) and spontaneous vesicular release long before they have completed axon development, which suggested a role for early neuronal activity in axon development and guidance (Catalano and Shatz, 1998). Cell-autonomous silencing of neurons in vivo by transfection of the hyperpolarizing inward-rectifying potassium channel Kir2.1 in olfactory neurons (Yu et al., 2004), in RGCs (Hua et al., 2005) or in cortical pyramidal neurons (Mizuno et al., 2007; Wang et al., 2007), or in vitro through infusion of tetrodotoxin (which blocks action potentials generation) in co-cultures of thalamo-cortical projecting neurons (Uesaka et al., 2007) results in a decrease in terminal axon branching, indicating that synaptic activity is required for axons to fully develop their branching pattern. Moreover, inhibition of synaptic release by expression of tetanus toxin light chain (TeTN-LC; Wang et al., 2007) also abolished terminal axon branching, suggesting that the formation of functional presynaptic release sites is required cell autonomously to control terminal axon branching. However, one potential limitation of the experiments involving TeTN-LC is that it blocks most VAMP-mediated vesicular release (with the exception of VAMP7, also called tetanus toxin–independent VAMP, or TI-VAMP). Therefore TeTN-LC action may not be limited to blocking synaptic vesicle release, but could also inhibit peptide release through vesicles containing neurotrophins for example, or other important trophic factors required for axon branching. More recent experiments through silencing of postsynaptic targets revealed that branching of callosal or thalamocortical axons is also dependent upon the activity of the postsynaptic targets (Mizuno et al., 2010; Yamada et al., 2010), albeit activity of the presynaptic neuron is required earlier during the branching process than activity of the postsynaptic targets (Mizuno et al., 2010). Activity is also required in some neurons at the phase of axon elongation through a feedback loop involving the activity-dependent up-regulation of guidance molecules (Mire et al., 2012).How much does spontaneous or evoked neuronal activity contribute to branching? Reduction of neuronal activity through hyperpolarization induced by overexpression of Kir2.1 significantly reduces axon branching without completely eliminating it (Hua et al., 2005; Mizuno et al., 2007; Wang et al., 2007). Activity seems to serve as a competitive regulator of axon branching with regard to its neighbors because silencing of neighboring axons restores normal branching (Hua et al., 2005). Interestingly, neuronal activity induces neurotrophin expression locally, suggesting that activity can contribute to branching partly through activation of activity-independent branching mechanisms (Calinescu et al., 2011).Neuronal activity can regulate branching through modification of the actin cytoskeleton via RhoA activation (Ohnami et al., 2008), and mRNA accumulates at presynaptic sites, indicating a correlation between local translation and synaptic activity (Lyles et al., 2006; Taylor et al., 2013). Neuronal activity is associated with changes in intracellular Ca2+ signaling, which has been shown to play a deterministic function in axon growth (Gomez and Spitzer, 1999). Calcium signaling activates the Ca2+/calmodulin-dependent kinases (CAMKs) that are known to regulate axon branching in vitro (Wayman et al., 2004; Ageta-Ishihara et al., 2009) and in vivo (Ageta-Ishihara et al., 2009).

Stabilization and refinement of the axonal arborization.

Axon branches are often formed in excess during development, then later refined to select for specific neural circuits (Luo and O’Leary, 2005). Long-range axon branch retraction has long been observed in layer V cortical neurons that initially project to the midbrain, hindbrain, and spinal cord (O’Leary and Terashima, 1988; Bastmeyer and O’Leary, 1996). At later stages, pyramidal neurons from the primary visual cortex will retract their spinal projection through axon pruning, whereas pyramidal neurons from the primary motor cortex will stabilize this projection but retract their axonal branches growing toward visual targets such as the superior colliculus. The molecular mechanisms controlling this area-specific pattern of axon branch pruning are still poorly understood, but seem to involve extracellular cues such as semaphorins and Rac1-dependent signaling (Bagri et al., 2003; Low et al., 2008; Riccomagno et al., 2012). Another example is the well-characterized refinement of retino-geniculate axons during the selective elimination of binocular input of RGC axon synapses onto relay neurons in the dorsal lateral geniculate nucleus (Muir-Robinson et al., 2002). Interestingly, some axons use caspase-dependent pathways locally to induce the selective retraction of axon branches during the process of pruning (Nikolaev et al., 2009; Simon et al., 2012).Circuit refinement and selective branch retraction can be observed in vivo at the level of the neuromuscular junction where individual branches of motor axons are eliminated asynchronously (Keller-Peck et al., 2001). In the developing CNS, neurotrophin-induced branch retraction can also be observed in a context of competition between neighboring axons (Singh et al., 2008). One other way of stabilizing axon branches is through the formation of synapses with postsynaptic targets. In the visual system, the initial axon arbor is refined to establish ocular dominance through activity-dependent retraction of less active branches (Ruthazer et al., 2003). Time-lapse imaging of RGC axons in zebrafish or in Xenopus tadpole revealed that the formation of presynaptic sites occurs concomitantly to axon branching, and branches that form presynaptic structures are less likely to retract (Meyer and Smith, 2006; Ruthazer et al., 2006). The stabilization of axon branches through formation of synaptic contacts parallels with the stabilization of dendritic branches through synapse formation and stabilization (Niell et al., 2004; J. Li et al., 2011). The role of presynaptic bouton formation goes beyond the stabilization of axonal branches because in vivo, new axon branches can emerge from existing presynaptic terminals (Alsina et al., 2001; Javaherian and Cline, 2005; Panzer et al., 2006).In conclusion, axon growth and branching can be regulated by both activity-dependent and activity-independent mechanisms during development. However, for mammalian CNS axons, much more work is needed to define (1) the precise molecular mechanisms underlying axon branching; (2) the cellular and molecular mechanisms regulating the key transition between axon growth and branching when axons start forming presynaptic contacts with their postsynaptic partners; and (3) the mechanisms regulating axon pruning during synapse elimination.  相似文献   

19.
20.
Planar cell polarity (PCP) refers to the coordinated alignment of cell polarity across the tissue plane. Key to the establishment of PCP is asymmetric partitioning of cortical PCP components and intercellular communication to coordinate polarity between neighboring cells. Recent progress has been made toward understanding how protein transport, endocytosis, and intercellular interactions contribute to asymmetric PCP protein localization. Additionally, the functions of gradients and mechanical forces as global cues that bias PCP orientation are beginning to be elucidated. Together, these findings are shedding light on how global cues integrate with local cell interactions to organize cellular polarity at the tissue level.The collective alignment of cell polarity across the tissue plane is a phenomenon known as planar cell polarity (PCP). Exemplified by the uniform orientation of bristles covering the insect epidermis or of the hairs covering the mammalian body surface (Fig. 1 A), PCP patterns can align over thousands, even billions of cells. This phenomenon is controlled by the so-called PCP pathway, which integrates global directional cues to produce locally polarized cell behaviors. There has been a recent surge in interest in PCP after discoveries that various processes such as vertebrate gastrulation, mammalian ear patterning and hearing, and neural tube closure all require a conserved set of PCP genes (Heisenberg et al., 2000; Tada and Smith, 2000; Wallingford et al., 2000; Kibar et al., 2001; Murdoch et al., 2001; Curtin et al., 2003; Montcouquiol et al., 2003; Copley et al., 2013). Since that time, the PCP pathway has been found to coordinate cell behaviors in numerous diverse settings including polarized ciliary beating in the trachea and brain ventricles (Tissir et al., 2010; Vladar et al., 2012), oriented cell divisions (Gong et al., 2004; Baena-López et al., 2005; Ségalen et al., 2010; Mao et al., 2011), lung branching (Yates et al., 2010), and hair follicle alignment (Guo et al., 2004; Devenport and Fuchs, 2008), to name a few (Fig. 1). Genetic disruptions to PCP cause severe developmental abnormalities in vertebrates, notably neural tube defects, left/right patterning defects, and ciliopathies, which highlights the essential requirement for PCP in development (Kibar et al., 2001; Murdoch et al., 2001; Curtin et al., 2003; Wang et al., 2006a,b; Kim et al., 2010; Song et al., 2010).Open in a separate windowFigure 1.Planar cell polarity and the core PCP components. (A and B) The Drosophila wing blade and mammalian epidermis illustrate the phenomenon of PCP. In both cases, hairs point in a single direction along the tissue axis, where they align locally with neighboring hairs and globally across the tissue. Whereas Drosophila wing hairs are produced by single cells, mammalian hairs emerge from multicellular hair follicles, which orient as a unit. A conserved PCP pathway controls the collective alignment of both types of structures. (C) Core PCP components localize to the plasma membrane and asymmetrically segregate along the epithelial plane as indicated.Like many types of cell polarity, the establishment of PCP involves (1) a global orienting cue, (2) asymmetric segregation of dedicated polarity proteins, and (3) translation of polarity information into polarized outputs. But unlike other types of cell polarity, the PCP mechanisms we currently understand involve coupling between adjacent cells, allowing for the alignment of polarity over many cell distances.First described in insects and then genetically dissected in Drosophila melanogaster, PCP was long confined to the realm of experimental embryology and genetics until the discovery that the protein products of several PCP genes were localized asymmetrically within the cell, thrusting PCP into the domain of cell biology (for review see Strutt and Strutt, 2009). The challenge to understanding PCP on a molecular level is that long-range PCP is, in essence, an in vivo phenomenon that is difficult to recapitulate in a tissue culture dish. However, recent advances in imaging technology combined with increasingly sophisticated genetic tools are helping us to decipher the in vivo cell biology of PCP. In this review, I highlight some of the recent advances made toward understanding the cell biology underlying the establishment of coordinated polarized cell behaviors. For clarity, I limit discussions of PCP phenomena that meet the definition of PCP proposed by Goodrich and Strutt (2011): namely, that “cell–cell communication causes two or more cells to adopt coordinated polarity” in a process that is mechanistically “dependent upon planar polarity proteins.” Other aligned cellular patterns or examples of noncanonical Wnt signaling, sometimes described as “Wnt/PCP” signaling, will not be discussed.

PCP components and molecular asymmetries

Two molecular systems control PCP behavior, the “core” and the “Fat–Dachsous (Ft–Ds)” PCP pathways. A key feature of both is the asymmetric distribution of their constituents (Fig. 2). The core PCP pathway is composed of the multipass transmembrane proteins Frizzled (Fz), Van Gogh (Vang; also known as Strabismus/Stbm), and Flamingo (Fmi; also known as Starry night/Stan), and the cytosolic components Dishevelled (Dsh), Prickle (Pk), and Diego (Dgo). On one edge of the cell reside Fz, Dsh, and Dgo, and on the opposite side lie Vang and Pk (Figs. 1 C and 2 B; Axelrod, 2001; Strutt, 2001; Feiguin et al., 2001; Tree et al., 2002; Bastock et al., 2003). The atypical cadherin, Fmi, resides on both sides, where it forms homodimers between neighboring cells (Usui et al., 1999; Shimada et al., 2001). These molecular asymmetries are observed in sensory hair cells of the vertebrate inner ear (Wang et al., 2005, 2006a,b; Montcouquiol et al., 2006; Deans et al., 2007; Song et al., 2010), the mammalian epidermis (Devenport and Fuchs, 2008; Devenport et al., 2011), brain ventricles (Tissir et al., 2010), and trachea (Vladar et al., 2012). Mutations in core PCP components lead to a loss or randomization of polarity and misalignment of cellular structures along the tissues axis.Open in a separate windowFigure 2.Asymmetric localization of PCP components. (A) PCP asymmetry develops progressively from an initially uniform distribution of core PCP proteins. Fz, Dsh, and Dgo (red) localize to the distal/posterior edge, whereas Vang and Pk (turquoise) localize to the proximal/anterior side. Fmi (dark blue) localizes to both sides, where it forms homodimers between neighboring cells. (B) Feedback interactions between core PCP components. A Fz–Fmi complex interacts preferentially with a Vang–Fmi complex between cells, whereas proximal and distal complexes antagonize one another within the cell. (C) Ds and Fj are expressed in opposing gradients in the Drosophila wing blade. Fj positively modulates Ft activity, leading to a gradient of Ft activity across the wing (not depicted). (D) Graded expression of Ds and Fj leads to asymmetric cellular localization of Ds and Ft, which form heterodimers between adjacent cells. Dachs, a downstream component of the Ft–Ds pathway, also localizes asymmetrically in association with Ds.The Ft–Ds pathway includes the large protocadherins Ft and Ds and the Golgi resident transmembrane kinase, Four-jointed (Fj; for review see Matis and Axelrod, 2013; Thomas and Strutt, 2012). Similar to the core system, Ft–Ds also displays molecular asymmetries in flies. Ds and its ligand Ft accumulate on opposite cell edges, where they form intercellular heterophilic interactions (Fig. 2 D; Matakatsu and Blair 2004; Ambegaonkar et al., 2012; Brittle et al., 2012). Unlike the core components, Ds and Fj are expressed in complementary gradients in the Drosophila eye and developing wing, which contribute to the cellular asymmetries of Ds and Ft (Fig. 2, C and D). Whether Ft–Ds–Fj gradients and asymmetries are conserved in vertebrate systems has yet to be determined.

Segregation of cortical polarity proteins: Shaking hands with the enemy

The asymmetric segregation of Fz–Dsh–Fmi and Vang–Pk–Fmi complexes to opposite sides of the cell relies on their mutual exclusion intracellularly and their preferential binding between neighboring cells (Fig. 2 B; for review see Strutt and Strutt, 2009). There is mutual interdependence among core PCP components for their asymmetric localization. Depletion of any one core PCP component results in a loss of asymmetry of all the others. In addition, PCP asymmetry develops progressively from initially uniform distributions (Fig. 2 A). Thus, PCP asymmetry can be thought of not as a simple hierarchy of interactions, but the result of feedback amplification of an initial directional bias.

Intercellular interactions.

PCP requires cell–cell communication, mediated by the transmembrane components of the core system, where it is thought that Fz–Fmi on one cell interacts with Vang–Fmi on its neighbor. These interactions are best understood in the Drosophila wing blade, where PCP controls the alignment of wing hairs along the proximal–distal axis (Figs. 1 A and 2, A and B). In the wing, Vang–Pk localize to the proximal face of each cell, whereas Fz–Dsh–Dgo localize distally adjacent to the wing hair (Figs. 1 C and 2, A and B; Axelrod, 2001; Strutt, 2001; Tree et al., 2002; Bastock et al., 2003; Das et al., 2004). By generating mutant clones and examining PCP localization at the clone border, the intercellular interactions between neighboring cells can be assessed in vivo. For example, when Fz is lacking within a clone, leaving only Vang–Fmi available at cell junctions, then Fz–Fmi in adjacent wild-type cells is recruited to the clone border (Chen et al., 2008). Vang mutant clones produce a similar effect, but in this case the excess Fz recruits Vang to clone borders (Bastock et al., 2003). What mediates these intercellular asymmetric interactions? One possibility is that Vang and Fz interact directly, and in vitro binding assays between the Fz extracellular domain and Vang suggest that this mechanism is possible (Wu and Mlodzik, 2008). However, mutants of Fz or Vang lacking their extracellular domains can still recruit one another between cells, which suggests that something else must bridge the two proteins (Chen et al., 2008). The seven-pass transmembrane cadherin, Fmi, likely performs this function. Fmi is essential for the junctional recruitment of Fz and Vang, and Fmi homodimers appear to be functionally asymmetric (Chen et al., 2008; Strutt and Strutt, 2008; Struhl et al., 2012). Clonal overexpression of Fmi preferentially recruits Fz to the clone border, even in the absence of Vang, which suggests that excess or unpaired Fmi is in a configuration that has higher affinity for Fmi–Fz than Fmi–Vang (Chen et al., 2008; Strutt and Strutt, 2008). Thus, Fmi may exist in two forms depending on whether it is paired with Fz or Vang, but the molecular basis for this difference is not known (Chen et al., 2008; Strutt and Strutt, 2008; Struhl et al., 2012).

Amplification of asymmetry.

Intercellular Fz–Fmi and Vang–Fmi complexes can form between cells in any orientation, so how do they resolve into discrete and opposed asymmetric domains? One way is through clustering of Fz–Fmi and Vang–Fmi complexes of the same orientation, and the cytoplasmic PCP components are particularly important for this function. As PCP complexes grow increasingly asymmetric, they cluster into discrete puncta that are stably associated with the plasma membrane and are resistant to endocytosis (Strutt et al., 2011). FRAP analysis of Fz-containing puncta demonstrated that they are highly stable compared with diffuse Fz-GFP, and have limited lateral mobility within the membrane. In the absence of Dsh, Pk, or Dgo, the size, intensity, and stability of Fz-containing puncta are diminished (Strutt et al., 2011), whereas overexpression causes Fz accumulation and coalescence into larger puncta (Feiguin et al., 2001; Tree et al., 2002; Bastock et al., 2003). Although the precise mechanisms driving PCP puncta formation are not known, the cytoplasmic components do not affect endocytosis, which suggests that they contribute to puncta formation by clustering intercellular complexes (Strutt et al., 2011). Pk can interact homophilically (Jenny et al., 2003; Ayukawa et al., 2014), which might promote clustering of proximal Vang–Pk–Fmi complexes. It will also be interesting to determine whether the cytoskeleton is directly connected to PCP complexes to minimize lateral mobility within the membrane.A second mechanism contributing to PCP asymmetry is directed transport. Live imaging of fluorescently tagged PCP proteins in pupal wings showed that Fz- and Dsh-containing particles travel across the cell in a proximal-to-distal direction (Shimada et al., 2006; Matis et al., 2014; Olofsson et al., 2014). These particles most likely represent endosomes undergoing transcytosis, as they arise from the proximal cortex and are labeled by the endocytic tracer FM4-64. This mechanism could serve to amplify asymmetry or even provide the initial polarity bias by removing proximal Fz–Dsh–Fmi complexes and transporting them to the distal side. Directed PCP transport is mediated by an array of subapical, noncentrosomal microtubules (MTs) that align along the proximal–distal axis, with the plus ends oriented with a slight distal bias (Hannus et al., 2002; Shimada et al., 2006; Harumoto et al., 2010; Matis et al., 2014; Olofsson et al., 2014). Ft and Ds are required for proximal–distal MT alignment (Harumoto et al., 2010), which suggests that the Ft-Ds system may feed into the core PCP system by orienting cytoskeletal architecture to deliver Fz–Dsh–Fmi complexes to the distal edge of the cell.Directed transport of Vang-containing endosomes has not been reported in flies, but selective trafficking could target Vang to specific membrane domains. In mammalian cells, exit of the Vang homologue Vangl2 from the trans-Golgi network (TGN) requires Arfrp1 (an Arf-like GTPase) and the clathrin adaptor complex AP-1, neither of which are required for the transport of a mammalian Fz homologue Fz6 or Fmi/Celsr1, which suggests that the differential sorting of PCP complexes to opposite sides of the cell could initiate at TGN export (Guo et al., 2013). Whether newly synthesized Vang and Fz proteins are transported to opposing cell surfaces from the TGN has not yet been explored.Microtubule orientation also correlates with PCP asymmetry in mouse trachea epithelial cells, where PCP coordinates the alignment of motile cilia (Vladar et al., 2012). MTs are planar polarized with their plus ends oriented toward the Fz–Dvl domain, and disruption of MTs with nocodazole impairs core PCP localization. Similarly, MTs are needed to establish Pk asymmetry in gastrulating zebrafish embryos (Sepich et al., 2011). However, in the skin epithelium, MTs align perpendicular to the axis of PCP asymmetry (unpublished data). Thus, directed transport along MTs may not be required in all tissue types for the establishment of PCP asymmetry.

Negative regulation.

Repulsive interactions between Vang- and Fz-containing complexes may also contribute to the amplification of asymmetry, and cytoplasmic proteins have been proposed to perform this function. Pk and Dgo both bind to Dsh in vitro, interacting with the same domain on Dsh in a mutually exclusive manner (Jenny et al., 2005). In addition, overexpression of Pk can prevent Dsh translocation to the membrane (Tree et al., 2002; Carreira-Barbosa et al., 2003), which suggests that Pk binding to Dsh could displace it from the proximal side of the cell. On the distal side, Dgo binding to Dsh would prevent association with Pk, thus enhancing Dsh distal localization. This increase in Dsh and Pk asymmetry would then positively feed back by clustering the transmembrane components into stable membrane domains.Modulation of PCP protein levels by ubiquitin-mediated degradation also leads to feedback by restricting the amount of one PCP protein to antagonize another. In flies, regulation of Dsh by a Cullin-3-BTB E3 ubiquitin ligase complex limits its levels at cell junctions (Strutt et al., 2013a). Reduction of Cullin-3 leads to an increase in overall core PCP protein levels, a reduction of asymmetry, and defects in wing hair polarity, which is consistent with Dsh overexpression phenotypes (Strutt et al., 2013a). SkpA, a subunit of the SCF E3 ligase, regulates Pk levels by promoting its degradation in a Vang-dependent manner (Strutt et al., 2013b). In mice, Smurf E3 ligases ubiquitinate Pk and promote its local degradation by binding to phosphorylated Dvl2 (a mammalian homologue of Dsh; Narimatsu et al., 2009). Smurfs are required for Pk localization in the inner ear and floor plate, and their removal leads to defects in convergent extension (CE) and stereocilia alignment (Narimatsu et al., 2009). Thus, targeting Pk for degradation either balances total Pk protein levels or targets a specific pool of Pk for ubiquitination and proteasome degradation.

Tissue-level polarity cues: This way or that?

What provides the tissue-level polarity cue that biases core PCP asymmetry in one direction over another? This is perhaps the most fundamental, yet poorly understood, element of PCP. Current models propose that an upstream, graded cue provides an initial bias in PCP asymmetry by regulating the levels, localization, or activity of one or more of the core proteins. Gradients are attractive candidates for providing global polarity cues, as they can act across many cells and define the tissue boundaries over which polarity must be oriented.

Ft–Ds–Fj.

Unlike the core proteins, Ds and Fj are nonuniformly expressed in the Drosophila eye, wing, and abdominal segments, and as such, the Ft–Ds module has been proposed to provide a global polarity cue (Fig. 2, C and D; for review see Ma et al., 2003; Yang et al., 2002; Thomas and Strutt, 2012; Matis and Axelrod, 2013). Ft and Ds are heterodimeric cadherins, regulated by the Golgi kinase Fj (Ishikawa et al., 2008; Brittle et al., 2010; Simon et al., 2010). The complementary expression patterns of Ds and Fj are thought to give rise to asymmetric Ft and Ds protein localization, with Ft and Ds localizing to opposite sides of each cell (Fig. 2 D; Ambegaonkar et al., 2012; Bosveld et al., 2012; Brittle et al., 2012). Because Fj positively regulates the activity of Ft, a gradient of Ft activity is expressed across the wing complementary to that of its ligand, Ds (Simon et al., 2010). Mutations in the Ft–Ds system give rise to swirling wing hair patterns, and disrupt the global alignment of core PCP proteins, but not their asymmetric distributions.An appealing model for symmetry breaking in the early Drosophila wing is that cellular asymmetries of Ft–Ds polarize MT organization and promote the distal transport of Fz–Dsh–Fmi vesicles (Shimada et al., 2001, Harumoto et al., 2010; Matis and Axelrod, 2013; Matis et al., 2014). This would produce an initial bias in Fz–Dsh localization, which would then be amplified by feedback interactions. However, several pieces of evidence have prevented the model from gaining universal acceptance. First, Ds and Fj gradients are oriented in opposite directions with respect to the core PCP proteins in the wing compared with the eye and abdomen. This discrepancy has been rectified with the finding by two independent groups that cells interpret Ft–Ds–Fj gradients differently depending on which of two Pk isoforms is expressed (Ayukawa et al., 2014; Olofsson et al., 2014). Second, the Ft–Ds system can orient PCP independently of the core pathway, and thus the two systems orient polarity in parallel, as opposed to in a single, common pathway (Casal et al., 2006). Third, Ft–Ds mutations affect core PCP orientation only regionally in the wing, which suggests that, if Ft–Ds provides a global bias, other, redundant cues must also exist (Matakatsu and Blair, 2006; Matis et. al., 2014). Finally, the direction of Ft–Ds and core PCP asymmetry diverges late in wing development, where the two systems become completely uncoupled. Intriguingly, the extent of coupling depends on which isoform of Pk is expressed (Merkel et al., 2014). Perhaps the simplest explanation for Ft–Ds function is that it can both transmit polarity information independent of the core system and organize the cytoskeleton to provide an initial bias of core PCP asymmetry, but which mechanism predominates depends on the tissue and developmental stage.

Wnts.

Wnt proteins have long been considered attractive candidates to provide tissue-level polarity cues because Fz and Dsh are primary components of the Wnt–β-catenin signaling pathway. Wnts are secreted glycoproteins that bind to Fz and other receptors, and often display graded expression. In vertebrates, Wnts are clearly important regulators of PCP, but whether they act instructively or permissively is unclear. In zebrafish, Wnt5a and Wnt11 are required for CE movements during gastrulation, but uniform expression of Wnt11 rescues the mutant phenotype, which suggests that it is permissive rather than instructive (Heisenberg et al., 2000; Kilian et al., 2003). Wnt5a is expressed in a gradient along the axis of polarity in the mouse inner ear, where it interacts genetically with Vangl2 in cochlear hair cell orientation (Qian et al., 2007). In the mouse limb, Wnt5a and its atypical receptor Ror2 are required for limb elongation and the asymmetric localization of Vangl2 at the proximal face of converging and extending chondrocytes (Gao et al., 2011). Wnt5a is expressed in a distal-to-proximal gradient, which induces a gradient of Vangl2 phosphorylation. The functional consequences of Vangl2 phosphorylation are unknown but Vangl2 cellular asymmetry appears to be strongest distally, where Wnt5a and Vangl2 phosphorylation levels are highest (Gao et al., 2011).While several studies had argued against the involvement of Wnt proteins in Drosophila PCP (Lawrence et al., 2002; Chen et al., 2008), it was recently discovered that Wingless (Wg) and Wnt4a act redundantly to orient PCP in the wing, particularly near the wing margin (Wu et al., 2013). Misexpression of Wg or Wnt4a reorients wing hair polarity in a pattern reminiscent of Fz loss of function, which suggests that Wnt gradients may orient polarity by antagonizing Fz. Consistently, the ability of Fz and Vang to recruit one another between adjacent cells in culture was inhibited by the addition of Wg or Wnt4a, which suggests that Wnts could provide a polarizing cue by diminishing Fz–Vang interactions at the margin of the wing, where Wnt expression is highest (Wu et al., 2013). However, Wnt4a overexpression also reorients MT alignment, suggesting that Wnts may act as polarity cues thorough an effect on the cytoskeleton (Matis et. al., 2014). Alternatively, Sagner et al. (2012) suggest that Wg orients core PCP indirectly through its effects on wing patterning and growth. Although the evidence for Wnt gradients as global PCP cues is accumulating, the mechanisms by which they regulate core protein levels or activity remain to be elucidated.

Mechanical forces.

Anisotropic mechanical forces that accompany growth and morphogenesis can also provide global polarizing cues. During wing development, PCP reorients in response to extensive morphogenetic changes that elongate the wing along the proximal–distal axis. In early pupal wings, PCP aligns toward the wing margin and then reorients during wing elongation and contraction of the wing hinge (Aigouy et al., 2010). These morphogenetic changes have broad effects on cell behavior, inducing cell elongation, oriented divisions, and cell rearrangements with a concomitant reorientation of PCP. Severing the wing pouch from the hinge blocks cell flows and PCP reorientation, which suggests that the anisotropic tension from hinge contraction drives tissue flow and the reorientation of polarity (Aigouy et al., 2010). Although this model doesn’t explain what initially biases PCP, it does demonstrate how the morphogenetic processes that shape tissues can completely remodel global PCP alignment. This is an attractive model to explain how PCP aligns over very large tissues, like the mammalian skin, where hairs consistently reorient along regions of extensive tissue elongation such as the face, limbs, and ears.

Downstream effectors of PCP: Steering the wheel

If PCP is the cell’s compass, it is also the steering wheel, directing downstream, polarized cell behaviors in response to global directional cues. PCP can polarize a wide range of cell behaviors, which suggests that it can intersect with numerous downstream effectors. We focus here on three examples where the molecular mechanisms linking core PCP to their polarized outputs have recently been elucidated.

Distal positioning of wing hairs.

Each cell of the Drosophila wing blade emits a single actin-rich protrusion from its distal edge. The placement of the wing hair strongly correlates with the position of Fz–Dsh–Fmi, which suggests that core proteins may localize cytoskeletal regulators to distinct positions within the cell (Strutt and Warrington, 2008). On the proximal side, Vang recruits a group of proteins that negatively regulate actin prehair formation: Inturned, Fuzzy, and Fritz (Adler et al., 2004; Strutt and Warrington, 2008). These three proteins regulate Multiple Wing Hairs, a GTPase-binding/formin-homology 3 (GBD/FH3) domain protein thought to repress actin polymerization (Strutt and Warrington, 2008; Yan et al., 2008). This restricts actin nucleation to distal positions within the cell, and in the absence of Multiple Wing Hairs, ectopic actin bundles form across the apical surface (Wong and Adler, 1993). On the distal side, casein kinase 1 γ CK1g/gilamesh is required to further refine prehair nucleation to a single site through a parallel mechanism involving Rab11-dependent vesicle traffic to the site of prehair formation (Gault et al., 2012). Rho and Rho kinase (Drok) have also been implicated in wing hair formation, but their roles are difficult to dissect due to the numerous functions of Rho in cell shape and cell division (Winter et al., 2001; Yan et al., 2009).

Actomyosin contraction and convergent extension (CE).

CE was the first vertebrate process to be linked molecularly to PCP (Wallingford et al., 2000). During CE, mesenchymal cells elongate, form mediolateral-directed protrusions, and intercalate mediolaterally, narrowing the mediolateral axis while simultaneously lengthening the anterior–posterior (A-P) axis (Fig. 3 A; Keller, 2002). Mediolateral polarization, elongation, and intercalation are lost when core PCP components are disrupted, leading to a failure in CE (Tada and Smith, 2000; Wallingford et al., 2000; Goto and Keller, 2002; Jessen et al., 2002). While several PCP-dependent mechanisms have been proposed to mediate CE movements, two recent studies provide direct mechanistic links between asymmetrically localized core PCP components and CE behaviors. In neuroepithelial cells, PCP specifies the localization of myosin to the A-P faces of intercalating cells. Fmi/Celsr1 and Dvl recruit the formin DAAM1 to the A-P junction, which in turn binds and activates PDZ-RhoGEF. This likely activates RhoA and myosin contractility specifically at A-P junctions, resulting in medial-directed cell intercalation and neural plate bending (Nishimura et al., 2012). A similar mechanism was found to drive CE movements of mesenchymal cells during Xenopus gastrulation. In this case, Fritz and Dsh help to localize septins to mediolateral vertices, where they spatially restrict cortical actomyosin contractility and junctional shrinking to A-P cell edges, thus driving cell intercalation (Fig. 3 A; Kim et al., 2010; Shindo and Wallingford, 2014). Together these studies show how asymmetric PCP localization produces collectively polarized cell behaviors through spatial modulation of the cytoskeleton.Open in a separate windowFigure 3.Polarized cell behaviors controlled by PCP. (A) PCP drives convergent extension (CE). CE in vertebrates is driven by mediolateral intercalation, which narrows the mediolateral axis while simultaneously lengthening the A-P axis. Mediolateral intercalation is accompanied by cell polarization and elongation and the formation of mediolateral protrusions, all of which require core PCP function. Pk localizes anteriorly (Ciruna et. al., 2006; Yin et al., 2008), whereas Dsh localizes posteriorly (Yin et. al., 2008). In addition, PCP proteins recruit myosin to A-P cell borders, leading to actomyosin contractility and junctional shrinking. (B) Asymmetric cell division. Drosophila sensory organ precursors (SOPs) divide asymmetrically along the epithelial plane, giving rise to distinct anterior and posterior daughters. Spindle alignment along the A-P axis is PCP dependent. Dsh interacts with Mud/NuMA and the dynein complex posteriorly while Vang links Pins/LGN-Mud/NuMA-dynein on the anterior. This links astral MTs to the A-P cortex, bringing the spindle into register with the A-P axis. (C) Positioning of the kinocilium in the inner ear. The placement of kinocilium in sensory hair cells of the inner ear determines the position of V-shaped stereocilia bundles. Gαi and mPins/LGN localize on the abneural side on the hair cell, where they are required for abneural positioning the MT-based kinocilium. The collective alignment of kinocilia and stereocilia bundles across the epithelium requires the core PCP component Vangl2. Vangl2 (light green) localizes to the abneural side of supporting cells. Whether Fz (dark blue) associates on the opposite face is not yet clear (Ezan et. al., 2013).

Positioning of centrosomes and cilia.

PCP regulates the positioning of MT-based structures including the mitotic spindle and cilia. In Drosophila sensory organ precursors and early zebrafish embryos, PCP controls mitotic spindle orientation along the epithelial plane by interacting with the highly conserved spindle orientation complex, which links astral MTs to the cell cortex through Mud/NuMA-mediated recruitment of the dynein complex (Ségalen et al., 2010). To orient the spindle, posteriorly localized Dsh binds to Mud/NuMA, which recruits the dynein complex and astral MTs to the posterior cortex. On the anterior side, Pins/LGN recruits Mud/NuMA, bringing the spindle into A-P alignment (Fig. 3 B). Similarly, PCP was recently shown to interact with the spindle orientation machinery to position the kinocilium in nondividing cells of the inner ear (Ezan et al., 2013; Tarchini et al., 2013). In vestibular hair cells, Gαi and mPins/LGN localize to the abneural cortex, opposite Vangl2, where they are required for kinocilium positioning and subsequent alignment of stereocilia bundles (Fig. 3 C; Ezan et al., 2013). MT plus ends and dynein also show an abneural bias suggesting that Gαi-mPins/LGN induces pulling on MTs by a similar mechanism that orients the centrosome during spindle orientation. Vangl2 is required for the alignment of Gαi-Pins/LGN crescents between cells, coordinating kinocilia positioning and stereocilia polarity across the tissue (Fig. 3 C; Ezan et al., 2013). Thus the PCP pathway co-opts the spindle orientation machinery to specify not only the division plane but also cilia position in nondividing cells. As PCP is required for asymmetric cilia positioning in a wide range of cell types, including the node (Antic et al., 2010; Borovina et. al., 2010; Song et al., 2010; Hashimoto et. al., 2010), it will be interesting to determine whether this mechanism is conserved.

Concluding remarks

PCP is a fundamental and highly conserved process coordinating a vast number of polarized cell behaviors. While the number of functions ascribed to PCP continues to grow, an understanding of the mechanisms establishing PCP is still far from complete. The development of cellular asymmetry from uniform distributions is not well understood, and will benefit from recent advances in high-resolution, time-lapse imaging with photoconvertible fluorescent tags. Other important issues to resolve include deciphering the structural domains and biochemical interactions mediating intercellular communication, identifying the global cues that orient PCP especially in vertebrates, and deciphering the mechanisms by which complex multicellular structures, like lung branches and hair follicles, are oriented by the PCP machinery.  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