首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
Homologous recombination provides high-fidelity DNA repair throughout all domains of life. Live cell fluorescence microscopy offers the opportunity to image individual recombination events in real time providing insight into the in vivo biochemistry of the involved proteins and DNA molecules as well as the cellular organization of the process of homologous recombination. Herein we review the cell biological aspects of mitotic homologous recombination with a focus on Saccharomyces cerevisiae and mammalian cells, but will also draw on findings from other experimental systems. Key topics of this review include the stoichiometry and dynamics of recombination complexes in vivo, the choreography of assembly and disassembly of recombination proteins at sites of DNA damage, the mobilization of damaged DNA during homology search, and the functional compartmentalization of the nucleus with respect to capacity of homologous recombination.Homologous recombination (HR) is defined as the homology-directed exchange of genetic information between two DNA molecules (Fig. 1). Mitotic recombination is often initiated by single-stranded DNA (ssDNA), which can arise by several avenues (Mehta and Haber 2014). They include the processing of DNA double-strand breaks by 5′ to 3′ resection, during replication of damaged DNA, or during excision repair (Symington 2014). The ssDNA is bound by replication protein A (RPA) to control its accessibility to the Rad51 recombinase (Sung 1994, 1997a; Sugiyama et al. 1997; Morrical 2014). The barrier to Rad51-catalyzed recombination imposed by RPA can be overcome by a number of mediators, such as BRCA2 and Rad52, which serve to replace RPA with Rad51 on ssDNA, and the Rad51 paralogs Rad55-Rad57 (RAD51B-RAD51C-XRCC2-XRCC3) and the Psy3-Csm2-Shu1-Shu2 complex (SHU) (RAD51D-XRCC2-SWS1), which stabilize Rad51 filaments on ssDNA (see Sung 1997b; Sigurdsson et al. 2001; Martin et al. 2006; Bernstein et al. 2011; Liu et al. 2011; Qing et al. 2011; Amunugama et al. 2013; Zelensky et al. 2014). The Rad51 nucleoprotein filament catalyzes the invasion into a homologous duplex to produce a displacement loop (D-loop) (Fig. 1). At this stage, additional antirecombination functions are exerted by Srs2 (FBH1, PARI), which dissociates Rad51 filaments from ssDNA, and Mph1 (FANCM), which disassembles D-loops (see Daley et al. 2014). Upon Rad51-catalyzed strand invasion, the ATP-dependent DNA translocase Rad54 enables the invading 3′ end to be extended by DNA polymerases to copy genetic information from the intact duplex (Li and Heyer 2009). Ligation of the products often leads to joint molecules (JMs), such as single- or double-Holliday junctions (s/dHJs) or hemicatenanes (HCs), which must be processed to allow separation of the sister chromatids during mitosis. JMs can be dissolved by the Sgs1-Top3-Rmi1 complex (STR) (BTR, BLM-TOP3α-RMI1-RMI2) (see Bizard and Hickson 2014) or resolved by structure-selective nucleases, such as Mus81-Mms4 (MUS81-EME1), Slx1-Slx4, and Yen1 (GEN1) (see Wyatt and West 2014). Mitotic cells favor recombination events that lead to noncrossover events likely to avoid potentially detrimental consequences of loss of heterozygosity and translocations.Open in a separate windowFigure 1.Primary pathways for homology-dependent double-strand break (DSB) repair. Recombinational repair of a DSB is initiated by 5′ to 3′ resection of the DNA end(s). The resulting 3′ single-stranded end(s) invade an intact homologous duplex (in red) to prime DNA synthesis. For DSBs that are repaired by the classical double-strand break repair (DSBR) model, the displaced strand from the donor duplex pairs with the 3′ single-stranded DNA (ssDNA) tail at the other side of the break, which primes a second round of DNA synthesis. After ligation of the newly synthesized DNA to the resected 5′ strands, a double-Holliday junction (dHJ) intermediate is generated. The dHJ can be either dissolved by branch migration (indicated by arrows) into a hemicatenane (HC) leading to noncrossover (NCO) products or resolved by endonucleolytic cleavage (indicated by triangles) to produce NCO (positions 1, 2, 3, and 4) or CO (positions 1, 2, 5, and 6) products. Alternatively to the double-strand break repair (DSBR) pathway, the invading strand is often displaced after limited synthesis and the nascent complementary strand anneals with the 3′ single-stranded tail of the other end of the DSB. After fill-in synthesis and ligation, this pathway generates NCO products and is referred to as synthesis-dependent strand annealing (SDSA).

Table 1.

Evolutionary conservation of homologous recombination proteins between Saccharomyces cerevisiae and Homo sapiens
Functional classS. cerevisiaeH. sapiens
End resectionMre11-Rad50-Xrs2MRE11-RAD50-NBS1
Sae2CtIP
Exo1EXO1
Dna2-Sgs1-Top3-Rmi1DNA2-BLM-TOP3α-RMI1-RMI2
AdaptorsRad953BP1, MDC1
BRCA1
Checkpoint signalingTel1ATM
Mec1-Ddc2ATR-ATRIP
Rad53CHK2
Rad24-RFCRAD17-RFC
Ddc1-Mec3-Rad17RAD9-HUS1-RAD1
Dpb11TOPBP1
Single-stranded DNA bindingRfa1-Rfa2-Rfa3RPA1-RPA2-RPA3
Single-strand annealingRad52RAD52
Rad59
MediatorsBRCA2-PALB2
Rad52
Strand exchangeRad51RAD51
Rad54RAD54A, RAD54B
Rdh54
Rad51 paralogsRad55-Rad57RAD51B-RAD51C-RAD51D-XRCC2-XRCC3
Psy3-Csm2-Shu1-Shu2RAD51D-XRCC2-SWS1
AntirecombinasesSrs2FBH1, PARI
Mph1FANCM
RTEL
Resolvases and nucleasesMus81-Mms4MUS81-EME1
Slx1-Slx4SLX1-SLX4
Yen1GEN1
Rad1-Rad10XPF-ERCC1
DissolutionSgs1-Top3-Rmi1BLM-TOP3α-RMI1-RMI2
Open in a separate windowThe vast majority of cell biological studies of mitotic recombination in living cells are performed by tagging of proteins with genetically encoded green fluorescent protein (GFP) or similar molecules (Shaner et al. 2005; Silva et al. 2012). In this context, it is important to keep in mind that an estimated 13% of yeast proteins are functionally compromised by GFP tagging (Huh et al. 2003). By choosing fluorophores with specific photochemical properties, it has been possible to infer biochemical properties, such as diffusion rates, protein–protein interactions, protein turnover, and stoichiometry of protein complexes at the single-cell level. To visualize the location of specific loci within the nucleus, sequence-specific DNA-binding proteins such the Lac and Tet repressors have been used with great success. Specifically, tandem arrays of 100–300 copies of repressor binding sites are inserted within 10–20 kb of the locus of interest in cells expressing the GFP-tagged repressor (Straight et al. 1996; Michaelis et al. 1997). In wild-type budding yeast, such protein-bound arrays are overcome by the replication fork without a cell-cycle delay or checkpoint activation (Dubarry et al. 2011). However, the arrays are unstable in rrm3Δ and other mutants (Dubarry et al. 2011). More pronounced DNA replication blockage by artificial protein-bound DNA tandem arrays has be observed in fission yeast, which is accompanied by increased recombination and formation of DNA anaphase bridges (Sofueva et al. 2011). Likewise, an array of Lac repressor binding sites was reported to induce chromosomal fragility in mouse cells (Jacome and Fernandez-Capetillo 2011). However, these repressor-bound arrays generally appear as a focus with a size smaller than the diffraction limit of light, which is in the range 150–300 nm for wide-field light microscopy.  相似文献   

3.
Recent advances in the characterization of the archaeal DNA replication system together with comparative genomic analysis have led to the identification of several previously uncharacterized archaeal proteins involved in replication and currently reveal a nearly complete correspondence between the components of the archaeal and eukaryotic replication machineries. It can be inferred that the archaeal ancestor of eukaryotes and even the last common ancestor of all extant archaea possessed replication machineries that were comparable in complexity to the eukaryotic replication system. The eukaryotic replication system encompasses multiple paralogs of ancestral components such that heteromeric complexes in eukaryotes replace archaeal homomeric complexes, apparently along with subfunctionalization of the eukaryotic complex subunits. In the archaea, parallel, lineage-specific duplications of many genes encoding replication machinery components are detectable as well; most of these archaeal paralogs remain to be functionally characterized. The archaeal replication system shows remarkable plasticity whereby even some essential components such as DNA polymerase and single-stranded DNA-binding protein are displaced by unrelated proteins with analogous activities in some lineages.Double-stranded DNA is the molecule that carries genetic information in all cellular life-forms; thus, replication of this genetic material is a fundamental physiological process that requires high accuracy and efficiency (Kornberg and Baker 2005). The general mechanism and principles of DNA replication are common in all three domains of life—archaea, bacteria, and eukaryotes—and include recognition of defined origins, melting DNA with the aid of dedicated helicases, RNA priming by the dedicated primase, recruitment of DNA polymerases and processivity factors, replication fork formation, and simultaneous replication of leading and lagging strands, the latter via Okazaki fragments (Kornberg and Baker 2005; Barry and Bell 2006; Hamdan and Richardson 2009; Hamdan and van Oijen 2010). Thus, it was a major surprise when it became clear that the protein machineries responsible for this complex process are drastically different, especially in bacteria compared with archaea and eukarya. The core components of the bacterial replication systems, such as DNA polymerase, primase, and replication helicase, are unrelated or only distantly related to their counterparts in the archaeal/eukaryotic replication apparatus (Edgell 1997; Leipe et al. 1999).The existence of two distinct molecular machines for genome replication has raised obvious questions on the nature of the replication system in the last universal common ancestor (LUCA) of all extant cellular life-forms, and three groups of hypotheses have been proposed (Leipe et al. 1999; Forterre 2002; Koonin 2005, 2006, 2009; Glansdorff 2008; McGeoch and Bell 2008): (1) The replication systems in Bacteria and in the archaeo–eukaryotic lineage originated independently from an RNA-genome LUCA or from a noncellular ancestral state that encompassed a mix of genetic elements with diverse replication strategies and molecular machineries. (2) The LUCA was a typical cellular life-form that possessed either the archaeal or the bacterial replication apparatus in which several key components have been replaced in the other major cellular lineage. (3) The LUCA was a complex cellular life-form that possessed both replication systems, so that the differentiation of the bacterial and the archaeo–eukaryotic replication machineries occurred as a result of genome streamlining in both lines of descent that was accompanied by differential loss of components. With regard to the possible substitution of replication systems, a plausible mechanism could be replicon takeover (Forterre 2006; McGeoch and Bell 2008). Under the replicon takeover hypothesis, mobile elements introduce into cells a new replication system or its components, which can displace the original replication system through one or several instances of integration of the given element into the host genome accompanied by inactivation of the host replication genes and/or origins of replication. This scenario is compatible with the experimental results showing that DNA replication DNA in Escherichia coli with an inactivated DnaAgene or origin of replication can be rescued by the replication apparatus of R1 or F1 plasmids integrated into the bacterial chromosome (Bernander et al. 1991; Koppes 1992). Furthermore, genome analysis suggests frequent replicon fusion in archaea and bacteria (McGeoch and Bell 2008); in particular, such events are implied by the observation that in archaeal genomes, genes encoding multiple paralogs of the replication helicase MCM and origins of replication are associated with mobile elements (Robinson and Bell 2007; Krupovic et al. 2010). Replicon fusion also is a plausible path from a single origin of replication that is typical of bacteria to multiple origins present in archaea and eukaryotes. However, all the evidence in support of frequent replicon fusion and the plausibility of replicon takeover notwithstanding, there is no evidence of displacement of the bacterial replication apparatus with the archaeal version introduced by mobile elements, or vice versa, displacement of the archaeal machinery with the bacterial version, despite the rapid accumulation of diverse bacterial and archaeal genome sequences. Thus, the displacement scenarios of DNA replication machinery evolution are so far not supported by comparative genomic data.Regardless of the nature of the DNA replication system (if any) in the LUCA and the underlying causes of the archaeo–bacterial dichotomy of replication machineries, the similarity between the archaeal and eukaryotic replication systems is striking (Leipe et al. 1999; Bell and Dutta 2002; Bohlke et al. 2002; Kelman and White 2005; Barry and Bell 2006). Thus, the archaeal replication system appears to be an ancestral version of the eukaryotic system and hence a good model for functional and structural studies aimed at gaining mechanistic insights into eukaryotic replication.

Table 1.

The relationship between archaeal and eukaryotic replication systems
Archaea (projection for LACA)Eukaryotes (projection for LECA)Comments
ORC complex
arORC1Orc1, Cdc6In LACA the ORC/Cdc6 complex probably consisted of two distinct subunits, and in LECA of six distinct. Both complexes might possess additional Orc6 and Cdt1 components.
arORC2Orc2, Orc3, Orc4, Orc5
TFIIB or homologaOrc6
WhiP or other wHTH proteinaCdt1
CMG complex
Archaeal Cdc45/RecJCdc45In many archaea and eukaryotes, CDC45/RecJ apparently contain inactive DHH phosphoesterase domains.
The RecJ family is triplicated in euryarchaea, and some of the paralogs could be involved in repair.
MCM is independently duplicated in several lineages of euryarchaea.
McmMcm2, Mcm3, Mcm4, Mcm5, Mcm6, Mcm7
Gins23Gins2, Gins3
Gins15Gins1, Gins5
Inactivated MCM homologaMcm10
CMG activation factors
RecQ/Sld2There is no evidence that kinases and phosphatases in archaea are directly involved in replication, although they probably regulate cell division.
Treslin/Sld3
TopBP1/Dpb11
STKCDK, DDK
PP2CPP2C
Primases
Prim1/p48PriSIn eukaryotes, Pol α is involved in priming by adding short DNA fragments to RNA primers.
In archaea, DnaG might be involved in priming specifically on the lagging strand.
Prim2a/p58PriL
DnaG
Polymerases
PolB3Pol α, Pol δ, Pol ζNo eukaryotic homologs of DP2 are known, but Zn fingers of Pol ε are apparently derived from DP2.
PolB1Pol ε
DP1B subunits of Pol α, Pol δ, Pol ζ, Pol ε
DP2
DNA polymerase sliding clamp and clamp loader
RFCLRFC1Eukaryotes have additional duplications of both RFCs and PCNA involved in checkpoint complexes (Rad27 and Rad1, Rad9, Hus1, respectively).
RFCSRFC2, RFC3, RFC4, RFC4
PCNAPCNA
Primer removal and gap closure
RNase H2RNase IIThere is a triplication of ligases (LigI, LigIII, LigIV) in eukaryotes, but only LigI is directly involved in replication.
In a few Halobacteria, ATP-dependent ligase is replaced by NAD-dependent ligase.
Fen1Fen1/EXO1, Rad2, Rad27
Lig1Lig1
SSB
arRPA1_longRpa1In Thermoproteales, RPA is displaced by the non-homologous ThermoSSB; two short RPA forms in many euryarchaea; expansion of short RPA forms in Halobacteria.
arRPA1_short and RPA2Rpa2
arCOG05741aRpa3
Open in a separate windowFor eukaryotic genes in Homo sapiens and Saccharomyces cerevisiae, gene names are indicated. Archaeal genes are denoted as in Barry and Bell (2006) or as introduced here.aNot confidently traced to LACA.In the last few years, there has been substantial progress in the study of the archaeal replication systems that has led to an apparently complete delineation of all proteins that are essential for replication (Berquist et al. 2007; Beattie and Bell 2011a; MacNeill 2011). The combination of experimental, structural, and bioinformatics studies has led to the discovery of archaeal homologs (orthologs) for several components of the replication system that have been previously deemed specific for eukaryotes (Barry and Bell 2006; MacNeill 2010, 2011; Makarova et al. 2012). Furthermore, complex evolutionary events that involve multiple lineage-specific duplications, domain rearrangements, and gene loss, and in part seem to parallel the evolution of the evolution of the replication system in eukaryotes, have been delineated for a variety of replication proteins in several archaeal lineages (Tahirov et al. 2009; Chia et al. 2010; Krupovic et al. 2010). Here we summarize these findings and present several additional case studies that show the complexity of evolutionary scenarios for the components of the archaeal replication machinery and new aspects of their relationship with the eukaryotic replication system.  相似文献   

4.
5.
6.
Two decades after the discovery that neural stem cells (NSCs) populate some regions of the mammalian central nervous system (CNS), deep knowledge has been accumulated on their capacity to generate new neurons in the adult brain. This constitutive adult neurogenesis occurs throughout life primarily within remnants of the embryonic germinal layers known as “neurogenic sites.” Nevertheless, some processes of neurogliogenesis also occur in the CNS parenchyma commonly considered as “nonneurogenic.” This “noncanonical” cell genesis has been the object of many claims, some of which turned out to be not true. Indeed, it is often an “incomplete” process as to its final outcome, heterogeneous by several measures, including regional location, progenitor identity, and fate of the progeny. These aspects also strictly depend on the animal species, suggesting that persistent neurogenic processes have uniquely adapted to the brain anatomy of different mammals. Whereas some examples of noncanonical neurogenesis are strictly parenchymal, others also show stem cell niche-like features and a strong link with the ventricular cavities. This work will review results obtained in a research field that expanded from classic neurogenesis studies involving a variety of areas of the CNS outside of the subventricular zone (SVZ) and subgranular zone (SGZ). It will be highlighted how knowledge concerning noncanonical neurogenic areas is still incomplete owing to its regional and species-specific heterogeneity, and to objective difficulties still hampering its full identification and characterization.The central nervous system (CNS) of adult mammals is assembled during developmental neurogenesis, and its architectural specificity is maintained through a vast cohort of membrane-bound and extracellular matrix molecules (Gumbiner 1996; Bonfanti 2006). Although CNS structure is sculpted by experience-dependent synaptic plasticity at different postnatal developmental stages (critical periods) (see Sale et al. 2009) and, to a lesser extent, during adulthood (Holtmaat and Svoboda 2009), the neural networks are rather stabilized in the “mature” nervous tissue (Spolidoro et al. 2009). The differentiated cellular elements forming adult neural circuitries remain substantially unchanged in terms of their number and types, because cell renewal/addition in the CNS is very low. This situation is intuitive because connectional, neurochemical, and functional specificities are fundamental features of the mature CNS in highly complex brains, allowing specific cell types to be connected and to act in a relatively invariant way (Frotscher 1992).Since the discovery of neural stem cells (NSCs) (Reynolds and Weiss 1992), we realized that the aforementioned rules of CNS stability have a main exception in two brain regions: the forebrain subventricular zone (SVZ) (Lois and Alvarez-Buylla 1994) and the hippocampal subgranular zone (SGZ) (Gage 2000). These “adult neurogenic sites” are remnants of the embryonic germinal layers (although indirectly for the SGZ, which forms ectopically from the embryonic germinative matrix), which retain stem/progenitor cells within a special microenvironment, a “niche,” allowing and regulating NSC activity (Kriegstein and Alvarez-Buylla 2009). In addition, the areas of destination (olfactory bulb and dentate gyrus) reached by neuroblasts generated within these neurogenic sites harbor specific, not fully identified yet, environmental signals allowing the integration of young, newborn neurons. These two “canonical” sites of adult neurogenesis have been found in all animal species studied so far, including humans (reviewed in Lindsey and Tropepe 2006; Bonfanti and Ponti 2008; Kempermann 2012; Grandel and Brand 2013). Although in several classes of vertebrates including fish, amphibians, and reptiles, adult neurogenesis is widespread in many areas of the CNS (Zupanc 2006; Chapouton et al. 2007; Grandel and Brand 2013), in mammals, the vast majority of the brain and spinal cord regions out of the germinal-layer-derived neurogenic sites are commonly referred to as “nonneurogenic parenchyma” (Sohur et al. 2006; Bonfanti and Peretto 2011; Bonfanti and Nacher 2012). However, this viewpoint has changed during the last few years. New examples of cell genesis, involving both neurogenesis and gliogenesis, have been shown to occur in the so-called nonneurogenic regions of the mammalian CNS (Horner et al. 2000; Dayer et al. 2005; Kokoeva et al. 2005; Luzzati et al. 2006; Ponti et al. 2008; reviewed in Butt et al. 2005; Nishiyama et al. 2009; Migaud et al. 2010; Bonfanti and Peretto 2011), suggesting that structural plasticity involving de novo neural cell genesis could be more widespread than previously thought. Apart from their temporal persistence (some of them represent examples of delayed developmental neurogenesis, which persist postnatally; see below), neurogliogenic processes vary as to their regional localization, origin, and final outcome. In this review, “noncanonical” neurogenic processes occurring in adult mammals will be reviewed by underlining their heterogeneity across the species and their differences in intensity and outcome with respect to canonical neurogenic sites.

Table 1.

Main sites of noncanonical neurogenesis in the mammalian brain
RatsMiceRabbitsMonkeys
NeocortexGould et al. 2001
Dayer et al. 2005a
Tamura et al. 2007
Shapiro et al. 2009Gould et al. 1999, 2001
Bernier et al. 2002
Nakatomi et al. 2002a
Pencea et al. 2001
Ohira et al. 2010a
Magavi et al. 2000a
Chen et al. 2004a
Vessal and Darian-Smith 2010a
Corpus callosumPencea et al. 2001
Piriform cortexbPekcec et al. 2006Shapiro et al. 2007Bernier et al. 2002
Olfactory tubercleShapiro et al. 2009Bedard et al. 2002b
StriatumDayer et al. 2005aShapiro et al. 2009Luzzati et al. 2006aBedard et al. 2002a;
2006a
Arvidsson et al. 2002a
Pencea et al. 2001
Liu et al. 2009a
Goldowitz and Hamre 1998a
Cho et al. 2007a
SeptumPencea et al. 2001
AmygdalaShapiro et al. 2009Luzzati et al. 2006aBernier et al. 2002
Hippocampus (Ammon’s horn)Rietze et al. 2000
Nakatomi et al. 2002a
ThalamusPencea et al. 2001
HypothalamusXu et al. 2005Kokoeva et al. 2007
Xu et al. 2005a
Pencea et al. 2001
Matsuzaki et al. 2009
Perez-Martin et al. 2010
Kokoeva et al. 2005a
Pierce and Xu 2010
Substantia nigraZhao et al. 2003
Zhao and Janson Lang 2009
Zhao et al. 2003
CerebellumPonti et al. 2008a
Brain stemBauer et al. 2005
Bauer et al. 2005
Open in a separate windowUnshaded rows, spontaneous (constitutive) neurogenesis; shaded rows, experimentally induced neurogenesis (growth factor infusion, lesion, etc.). No functional integration has been shown to occur in any of the studies reported here.aNeuronal differentiation of newborn cells has been well documented; in all other cases, neurogenesis has been shown only until the cell-specification step, and/or assessed with less accurate analyses (reslicing not performed, neuronal differentiation not clearly shown, very few cells shown in figures, insufficient or absent quantification).bNeurogenesis reported in this region has been denied by subsequent reports. Only a set of studies are reported; gliogenesis is not considered (data modified from Bonfanti and Peretto 2011).  相似文献   

7.
Dynamic changes in cytosolic and nuclear Ca2+ concentration are reported to play a critical regulatory role in different aspects of skeletal muscle development and differentiation. Here we review our current knowledge of the spatial dynamics of Ca2+ signals generated during muscle development in mouse, rat, and Xenopus myocytes in culture, in the exposed myotome of dissected Xenopus embryos, and in intact normally developing zebrafish. It is becoming clear that subcellular domains, either membrane-bound or otherwise, may have their own Ca2+ signaling signatures. Thus, to understand the roles played by myogenic Ca2+ signaling, we must consider: (1) the triggers and targets within these signaling domains; (2) interdomain signaling, and (3) how these Ca2+ signals integrate with other signaling networks involved in myogenesis. Imaging techniques that are currently available to provide direct visualization of these Ca2+ signals are also described.The recognition of Ca2+ as a key regulator of muscle contraction dates back to Sydney Ringer''s seminal observations in the latter part of the 19th Century (Ringer 1883; Ringer 1886; Ringer and Buxton 1887; see reviews by Martonosi 2000; Szent-Györgyi 2004). More recently, evidence is steadily accumulating to support the proposition that Ca2+ also plays a necessary and essential role in regulating embryonic muscle development and differentiation (Flucher and Andrews 1993; Ferrari et al. 1996; Lorenzon et al. 1997; Ferrari and Spitzer 1998, 1999; Wu et al. 2000; Powell et al. 2001; Jaimovich and Carrasco 2002; Li et al. 2004; Brennan et al. 2005; Harris et al. 2005; Campbell et al. 2006; Terry et al. 2006; Fujita et al. 2007; and see reviews by Berchtold et al. 2000; Ferrari et al. 2006; Al-Shanti and Stewart 2009). What is currently lacking, however, is extensive direct visualization of the spatial dynamics of the Ca2+ signals generated by developing and differentiating muscle cells. This is especially so concerning in situ studies. The object of this article, therefore, is to review and report the current state of our understanding concerning the spatial nature of Ca2+ signaling during embryonic muscle development, especially from an in vivo perspective, and to suggest possible directions for future research. The focus of our article is embryonic skeletal muscle development because of this being an area of significant current interest. Several of the basic observations reported, however, may also be common to cardiac muscle development and in some cases to smooth muscle development. What the recent development of reliable imaging techniques has most certainly done, is to add an extra dimension of complexity to understanding the roles played by Ca2+ signaling in skeletal muscle development. For example, it is clear that membrane-bound subcellular compartments, such as the nucleus (Jaimovich and Carrasco 2002), may have endogenous Ca2+ signaling activities, as do specific cytoplasmic domains, such as the subsarcolemmal space (Campbell et al. 2006). How these Ca2+ signals interact with specific down-stream targets within their particular domain, and how they might serve to communicate information among domains, will most certainly be one of the future challenges in elucidating the Ca2+-mediated regulation of muscle development.Any methodology used to study the properties of biological molecules and how they interact during development should ideally provide spatial information, because researchers increasingly need to integrate data about the interactions that underlie a biological process (such as differentiation) with information regarding the precise location within cells or an embryo where these interactions take place. Current Ca2+ imaging techniques are beginning to provide us with this spatial information, and are thus opening up exciting new avenues of investigation in our quest to understand the signaling pathways that regulate muscle development (
AnimalIntact animals/Cells in cultureCa2+ reporterReporter Loading ProtocolReference
Rat1° cultures prepared from hind limb muscle of neonatal rat pupsFluo 3-AMCells incubated in 5.4 µM reporter for 30 min at 25°C.Jaimovich et al. 2000
MouseMyotubes grown from C2C12 subclone of the C2 mouse muscle cell lineFluo 3-AMIncubated in 5 µM reporter plus 0.1% pluronic F-127 for 1 h at r.t.Flucher and Andrews 1993
Myotubes isolated from the intercostal muscles of E18 wild-type and RyR type 3-null mice.Fluo 3-AMCells incubated with 4 µM for 30 min at r.t.Conklin et al. 1999b
Myotubes in culture prepared from newborn mice.Fluo 3-AMCells incubated in 10 µM for 20 min.Shirokova et al. 1999
1° cultures prepared from hind limb muscle from newborn mice.Fluo 3-AMCells incubated in 5.4 µM reporter for 30 min at 25°C.Powell et al. 2001
Embryonic day 18 (E18) isolated diaphragm muscle fibersFluo 4-AMIncubated in 10 µM reporter for 30 min.Chun et al. 2003
ChickMyotubes prepared from leg or breast of 11-day chick embryosFluo 3-AMIncubated in 5 µM reporter plus 0.1% pluronic F-127 for 1 h at r.t.Flucher and Andrews 1993
Myoblasts isolated from thigh muscle of E12 embryos.Fluo 3-AM1 mM stock was diluted 1:200 with 0.2% pluronic F-127. Cells were incubated for 60 min at r.t. in the dark.Tabata et al. 2006
XenopusExposed myotome in dissected embryoFluo-3 AMIncubated dissected tissue in 10 µM reporter for 30–60 min.Ferrari and Spitzer 1999
1° myocyte cultures prepared from stage 15 Xenopus embryos.Fluo-4 AMCells incubated in 2 µM reporter plus 0.01% pluronic F-127 for 60 min.Campbell et al. 2006
ZebrafishIntact animalsCalcium green-1 dextran (10S)Reporter at 20 mM was injected into a single blastomere between the 32- and 128-cell stage.Zimprich et al. 1998
Intact animalsOregon Green 488 BAPTA dextranSingle blastomeres from 32-cell stage embryos injected with reporter (i.c. 100 µM) and tetramethylrhodamine dextran (i.c. 40 µM).Ashworth et al. 2001
Intact animalsOregon Green 488 BAPTA dextranMicroinjected with rhodamine dextran to give an intracellular concentration of ∼40 µM.Ashworth 2004
Intact animalsAequorinaEmbryos injected with 700 pg aeq-mRNA at the 1-cell stage and then incubated with 50 µM f-coelenterazine from the 64-cell stage.Cheung et al. 2006
Intact animalsAequorinTransgenic fish that express apoaequorin in the skeletal muscles were incubated with 50 µM f-coelenterazine from the 8-cell stage.Cheung et al. 2010
Open in a separate windowaExpression of aequorin was ubiquitous but it was suggested that the Ca2+ signals visualized in the trunk at the approximately 8–20-somite stage and at ∼47 hpf might play a role in muscle development.  相似文献   

8.
DNA Repair at Telomeres: Keeping the Ends Intact     
Christopher J. Webb  Yun Wu  Virginia A. Zakian 《Cold Spring Harbor perspectives in biology》2013,5(6)
  相似文献   

9.
Endocytosis: Past,Present, and Future     
Sandra L. Schmid  Alexander Sorkin  Marino Zerial 《Cold Spring Harbor perspectives in biology》2014,6(12)
Endocytosis may have been a driving force behind the evolution of eukaryotic cells. It plays critical roles in cell biology (e.g., signal transduction) and in organismal physiology (e.g., tissue morphogenesis).Endocytosis, the process of cellular ingestion, may have been the driving force behind evolution of the eucaryotic cell (de Duve 2007). Acquiring the ability to internalize macromolecules and digest them intracellularly would have allowed primordial cells to move out from their food sources and pursue a predatory existence; one that might have led to the development of endosymbiotic relationships with mitochondria and plastids. Thus, it is fitting that endocytosis was first discovered and named as the processes of cell “eating” and “drinking.” In 1883, the developmental biologist Ilya Metchnikoff coined the term phagocytosis, from the Greek “phagos” (to eat) and “cyte” (cell), after observing motile cells in transparent starfish larva surround and engulf small splinters that he had inserted (Tauber 2003). Decades later, in 1931, Warren H. Lewis, one of the earliest cell “cinematographers” coined the term pinocytosis, from the Greek “pinean” (to drink), after observing the uptake of surrounding media into large vesicles in cultured macrophages, sarcoma cells, and fibroblasts by time-lapse imaging (Lewis 1931; Corner 1967).Importantly, these pioneering studies also revealed that the function of endocytosis goes well beyond eating and drinking. Indeed, Metchnikoff, considered one of the founders of modern immunology, realized that the phagocytic behavior of the mesodermal amoeboid cells he had observed under the microscope could serve as a general defense system against invasive parasites, in the larva as in man. This revolutionary concept, termed the phagocytic theory, earned Metchnikoff the 1908 Nobel Prize in Physiology or Medicine for his work on phagocytic immunity, which he shared with Paul Ehrlich who discovered the complementary mechanisms of humoral immunity that led to the identification of antibodies (Vaughan 1965; Tauber 2003; Schmalstieg and Goldman 2008). The phagocytic theory was a milestone in immunology and cell biology, and formally gave birth to the field of endocytosis.Key discoveries over the intervening years, aided in large part by the advent of electron microscopy, revealed multiple pathways for endocytosis in mammalian cells that fulfill multiple critical cellular functions (Fig. 1). These mechanistically and morphologically distinct pathways, and their discoverers, include clathrin-mediated endocytosis (Roth and Porter 1964), caveolae uptake (Palade 1953; Yamada 1955), cholesterol-sensitive clathrin- and caveolae-independent pathways (Moya et al. 1985; Hansen et al. 1991; Lamaze et al. 2001), and, more recently, the large capacity CLIC/GEEC pathway (Kirkham et al. 2005). In place of Metchnikoff’s splinters, many of these discoveries resulted from the detection and tracking of internalized HRP-, ferritin-, or gold-conjugated ligands by electron microscopy. These electron-dense tracers allowed researchers to identify cellular structures associated with the uptake and intracellular sorting of receptor-bound ligands. A particularly striking example is the pioneering work of Roth and Porter, who in 1964 observed the uptake of yolk proteins into mosquito oocytes. To synchronize uptake, they killed female mosquitos at timed intervals after a blood feed and observed the sequential appearance of electron-dense yolk granules in coated pits, coated and uncoated vesicles, and progressively larger vesicles. Their remarkable observations accurately described coated vesicle budding, uncoating, homo- and heterotypic fusion events, as well as the emergence of tubules from early endosomes (Fig. 2), all of which are now known hallmarks of the early endocytic trafficking events.Open in a separate windowFigure 1.Time line for discoveries of endocytic pathways and their discoverers. Boxes are color-coded by pathway. *, Nobel laureate. HRP, horseradish peroxidase; CCVs, clathrin-coated vesicles; CCPs, clathrin-coated pits; EGFR, epidermal growth factor receptor; PM, plasma membrane; ER, endoplasmic reticulum; CLIC/GEEC, clathrin-independent carriers/GPI-enriched endocytic compartments.Open in a separate windowFigure 2.Fiftieth anniversary of the discovery of clathrin-mediated endocytosis by Roth and Porter (1964). The image is the hand-drawn summary of observations made by electron microscopic examination of the trafficking of yolk proteins in a mosquito oocyte. Note the many details, later confirmed and mechanistically studied over the intervening 50 years. These include the growth, invagination, and pinching off of coated pits (1,2), which concentrate cargo taken up by coated vesicles (3), the rapid uncoating of nascent-coated vesicles (4), homotypic fusion of nascent endocytic vesicles in the cell periphery (5), the formation of tubules from early endosomes (7), and changes in the intraluminal properties of larger endosomes (6). Finally, yolk proteins are stored in granules as crystalline bodies (8). (From Roth and Porter 1964; reprinted, with express permission, from Rockefeller University Press © 1964, The Journal of Cell Biology 20: 313–332, doi: 10.1083/jcb.20.2.313.)Another milestone in the field of endocytosis was the discovery of the lysosome by Christian de Duve (Appelmans et al. 1955). Whereas the finding of phagocytosis and other endocytic pathways was possible through microscopy, the discovery of lysosomes originated from a biochemical approach (Courtoy 2007), which benefited from the invention of the ultracentrifuge. de Duve and coworkers observed that preparations of acid phosphatase obtained by subcellular fractionation had an unusual behavior: contrary to most enzymatic activities, the activity of acid phosphatase increased rather than decreased with time, freezing–thawing of the fractions and the presence of detergents. Interestingly, the same was true for other hydrolases, which depended on acidic pH for their optimal activity. This led him to postulate that the acid hydrolases were contained in acidified membrane-bound vesicles. In collaboration with Alex Novikoff, he visualized these vesicles, the lysosomes, by electron microscopy (Beaufay et al. 1956) and later showed their content of acid phosphatase (Farquhar et al. 1972). In 1974, de Duve was awarded the Nobel Prize for Physiology or Medicine for his seminal finding of the lysosomes and peroxisomes. He shared it with Albert Claude and George E. Palade “for their discoveries concerning the structural and functional organization of the cell.” The importance of this work lies also in the significant therapeutic applications that followed. The discovery by Elizabeth Neufeld and collaborators of uptake of lysosomal enzymes by cells provided the foundation for enzyme replacement therapy for lysosomal storage disorders (Neufeld 2011).In the 1970s, research in endocytosis entered the molecular era. Using de Duve and Albert Claude-like methods of subcellular fractionation, Barbara M. Pearse purified clathrin-coated vesicles from pig brain (Pearse 1975). A year later, she isolated a major protein species of 180 kDa, which she named clathrin “to indicate the lattice-like structures which it forms” (Pearse 1976). It was a breakthrough that inaugurated the molecular dissection of clathrin-mediated endocytosis.Over the intervening years, the continued application of microscopy (which now spans from electron cryotomography to live cell, high-resolution fluorescence microscopy), genetics (in particular, in yeast, Caenorhabditis elegans and Drosophila melanogaster), biochemistry (including cell-free reconstitution of endocytic membrane trafficking events), as well as molecular and structural biology have revealed a great deal about the cellular machineries and mechanisms that govern trafficking along the endocytic pathway. A partial, and because of space limitations, necessarily incomplete list of milestones (YearMechanistic milestonesDiscoverers1973Identification of shibirets (dynamin) mutant in DrosophilaD. Suzuki and C. Poodry1974–1976Zipper mechanism for phagocytosisS. Silverstein1975–1976Isolation of CCVs, purification of clathrinB. Pearse1982–1984Phosphomannose, M6PR, and lysosomal targetingW. Sly, S. Kornfeld, E. Neufeld, G. Sahagian1983–1984Isolation of clathrin adapters/localization to distinct membranesJ. Keen, B. Pearse, M. Robinson1986Isolation of endocytosis mutants (End) in yeastH. Riezman1986–1987Isolation of vacuolar protein sorting mutants in yeastS. Emr, T. Stevens1986Endosome fusion in vitroJ. Gruenberg and K. Howell1986EGF and insulin receptor signaling from endosomesJ. Bergeron and B. Posner1986Macropinocytosis induced in stimulated cellsD. Bar-Sagi and J. Feramisco1987Endocytic sorting motifs (FxNPxY, YxxF)M. Brown and J. Goldstein, I. Trowbridge, T. McGraw1987–1989Cloning of CHC, CLC, AP2T. Kirchhausen, M. Robinson1988Isolation of biochemically distinct early and late endosomesS. Schmid and I. Mellman1989–1991Clathrin-mediated endocytosis reconstituted in vitroE. Smythe, G. Warren, S. Schmid1990Localization of endosomal Rab5 and Rab7P. Chavrier, R. Parton, M. Zerial1991Endosome to trans-Golgi network (TGN) transport reconstituted in vitroS. Pfeffer1992Rab5 and Rab4 as early endocytic regulators in vivoM. Zerial, R. Parton, I. Mellman1992–1995Caveolin/VIP21 identified as caveolar coat proteinR. Anderson, T. Kurzchalia, R. Parton, K. Simons1992Vacuolar fusion reconstituted in vitroW. Wickner1992–1994Trigger mechanism for phagocytosis of bacteriaS. Falkow, J. Galán, J. Swanson1993Actin’s role in endocytosis in yeastH. Riezman1993Isolation of autophagy mutants (Atg) in yeastY. Ohsumi1993PI3 kinase activity (PI3P) and endosome functionS. Emr1993Dynamin’s role in clathrin-mediated endocytosisR. Vallee, S. Schmid1995Dynamin assembles into ringsS. Schmid, P. De Camilli1996Clathrin-mediated endocytosis requirement for signalingS. Schmid1996Long distance retrograde transport of signaling endosomes in neuronsW. Mobley1996PI5 phosphatase activity (PI(4,5)P2) and clathrin-mediated endocytosisP. De Camilli1996Ubiquitin-dependent sorting in endocytosisR. Haguenauer-Tsapis; L. Hicke and H. Riezman1997AP3 and endosomal/lysosomal sortingJ. Bonifacino, S. Robinson1998FYVE fingers bind to PI3PH. Stenmark1998LBPA in MVB biogenesisT. Kobayashi, R. Parton, J. Gruenberg1997–1998Sorting nexinsG. Gill, S. Emr1998Structural basis for Y-based sorting signal recognitionD. Owen1998Retromer coat and endosome to TGN sortingS. Emr1998β-Propeller structure of clathrin heavy chain terminal domainT. Kirchhausen and S. Harrison1998Cargo-specific subpopulations of clathrin-coated pitsM. von Zastrow1999Structure of the clathrin coat protein superhelical motifsJ. Ybe and F. Brodsky1999Imaging green fluorescent protein–clathrin in living cellsJ. Keen1999Biochemical purification of Rab5 effectorsS. Christoforidis and M. Zerial1999Genetic screen for endocytosis mutants in C. elegansB. Grant2000Role of endocytosis in establishing morphogenic gradientsM. Gonzalez-Gaitan, S.M. Cohen2000Identification of GGA coats and lysosomal sortingJ. Bonifacino, S. Kornfeld, M. Robinson2000Identification of endosomal sorting complex required for transport (ESCRT) machinery for multivesicular body (MVB) formationS. Emr2001Ubiquitin-dependent sorting into MVBsR. Piper, S. Emr, H. Pelham2002Structure of the AP2 coreD. Owen2003Lipid conjugation of LC3/Atg8Y. Ohsumi2003–2004siRNA studies of endocytic componentsS. Robinson, E. Ungewickell, A. Sorkin2004BAR domains and membrane curvature generationH. McMahon, P. De Camilli20048-Å structure of a complete clathrin coatT. Kirchhausen and S. Harrison2005Modular design of yeast endocytosis machineryD. Drubin and M. Kaksonen2005Kinome-wide RNAi analysis of clathrin-mediated endocytosis (CME) and clathrin-independent endocytosis (CIE)M. Zerial and L. Pelkmans2006–2008Reconstitution of dynamin-mediated membrane fissionA. Roux, P. De Camilli, S. Schmid, J. Zimmerberg, V. Frolov2007Glycosphingolipid-induced endocytosisL. Johannes2009Reconstitution of Rab- and SNARE-dependent vacuolar and endosome fusion from purified componentsW. Wickner, M. Zerial2010Cavins as major caveolae coat componentsR. Parton; B. Nichols2010Reconstitution of ESCRT-dependent internal vesicle formationT. Wollert and J. Hurley2012Reconstitution of CCV formation from minimal componentsE. UngewickellOpen in a separate window  相似文献   

10.
The hippo pathway     
Harvey KF  Hariharan IK 《Cold Spring Harbor perspectives in biology》2012,4(8):a011288
  相似文献   

11.
Heparan Sulfate Proteoglycans     
Stephane Sarrazin  William C. Lamanna  Jeffrey D. Esko 《Cold Spring Harbor perspectives in biology》2011,3(7)
Heparan sulfate proteoglycans are found at the cell surface and in the extracellular matrix, where they interact with a plethora of ligands. Over the last decade, new insights have emerged regarding the mechanism and biological significance of these interactions. Here, we discuss changing views on the specificity of protein–heparan sulfate binding and the activity of HSPGs as receptors and coreceptors. Although few in number, heparan sulfate proteoglycans have profound effects at the cellular, tissue, and organismal level.Heparan sulfate proteoglycans (HSPGs) are glycoproteins, with the common characteristic of containing one or more covalently attached heparan sulfate (HS) chains, a type of glycosaminoglycan (GAG) (Esko et al. 2009). Cells elaborate a relatively small set of HSPGs (∼17) that fall into three groups according to their location: membrane HSPGs, such as syndecans and glycosylphosphatidylinositol-anchored proteoglycans (glypicans), the secreted extracellular matrix HSPGs (agrin, perlecan, type XVIII collagen), and the secretory vesicle proteoglycan, serglycin (Esko et al. 1985), which allowed functional studies in the context of a cell culture model (Zhang et al. 2006). A decade later, the first HSPG mutants in a model organism (Drosophila melanogaster) were identified (Rogalski et al. 1993; Nakato et al. 1995; Häcker et al. 1997; Bellaiche et al. 1998; Lin et al. 1999), which was followed by identification of mutants in nematodes, tree frogs, zebrafish, and mice (and3).3). HS is evolutionarily ancient and its composition has remained relatively constant from Hydra to humans (Yamada et al. 2007; Lawrence et al. 2008).

Table 1.

Heparan sulfate proteoglycans
ClassProteoglycanCore mass (kDa)aChain type (number)bTissueHuman disease
Membrane-boundSyndecan-1–syndecan-431–45HS (2–3) in Sdc2 and Sdc4; HS/CS (3–4 HS/1-2 CS) in Sdc1 and Sdc3Epithelial cells, fibroblasts
Glypican-1–glypican-657–69HS (1–3)Epithelial cells, fibroblastsSimpson–Golabi–Behmel syndrome (overgrowth) (GPC3) (Pilia et al. 1996); omodysplasia (skeletal dysplasia) (GPC6) (Campos-Xavier et al. 2009)
Betaglycan (part-time PG)110HS/CS (1–2)Fibroblasts
Neuropilin-1 (part-time PG)130HS or CS (1)Endothelial cells
CD44v337HS (1)Lymphocytes
Secretory vesiclesSerglycin10–19Heparin/CS (10–15)Mast cells, hematopoietic cells
Extracellular matrixPerlecan400HS (1–4)Basement membranesSchwartz–Jampel syndrome (skeletal dysplasia) (Nicole 2000; Arikawa-Hirasawa et al. 2001)
Agrin212HS (2–3)Basement membranes
Collagen XVIII150HS (1–3)Epithelial cells, basement membranesKnobloch syndrome type I (Sertie et al. 2000)
Open in a separate windowHS, heparan sulfate; CS, chondroitin sulfate; PG, proteoglycan.aThe variation in core mass is because of species differences.bThe number of chains is based on the number of putative attachment sites for chain initiation as well as data from the literature; the actual number of chains varies by method, tissue, and species.

Table 2.

Mutants altered in HSPG core proteins
GeneProteoglycanPhenotype (references)
Sdc1Syndecan-1Null allele: viable; increase in inflammation-mediated corneal angiogenesis (Gotte et al. 2002, 2005); corneal epithelial cells migrate more slowly, show reduced localization of α9 integrin during wound closure and fail to increase in proliferation after wounding (Stepp et al. 2002); enhanced leukocyte-endothelial interaction in the retina (Gotte et al. 2002, 2005); increase in medial and intimal smooth muscle cell replication and neointimal lesion after injury (Fukai et al. 2009); reduced cardiac fibrosis and dysfunction during angiotensin II–induced hypertension (Schellings et al. 2010); not required for follicle initiation and development (Richardson et al. 2009); accumulates plasma triglycerides, and shows prolonged circulation of injected human VLDL and intestinally derived chylomicrons (Stanford et al. 2009); juvenile mice resistant to carcinogen-induced tumorigenesis (McDermott et al. 2007); increased basal protein leakage and more susceptible to protein loss induced by combinations of IFN-γ, TNF-α, and increased venous pressure (Bode et al. 2008); exacerbates anti-GBM nephritis shifting Th1/Th2 balance toward a Th2 response (Rops et al. 2007); no role in hepatocyte infection by Plasmodium yoelii sporozoites (Bhanot 2002); normal larval development of Trichinella spiralis, but modestly reduced Th2 responses during infection (Beiting et al. 2006); less susceptible to Pseudomonas aeruginosa infection (Haynes et al. 2005); reduced P. aeruginosa infection rate and virulence (Park et al. 2001); protected from Staphylococcus aureus beta-toxin-induced lung injury (Hayashida et al. 2009a); exaggerated airway hyperresponsiveness, eosinophilia, and lung IL-4 responses to allergens (Xu et al. 2005); exaggerated CXC chemokines, neutrophilic inflammation, organ damage, and lethality in LPS endotoxemia (Hayashida et al. 2009b); prolonged recruitment of inflammatory cells in dextran sodium sulfate (DSS)-induced colitis and delayed type hypersensitivity (Masouleh et al. 2009; Floer et al. 2010).
Sdc2Syndecan-2No mutants reported. Sdc2 antisense impairs angiogenesis in human microvascular endothelial cells (Noguer et al. 2009); morpholinos inhibit cell migration and fibrillogenesis during embryogenesis in zebrafish (Arrington and Yost 2009).
Sdc3Syndecan-3Null allele: viable; altered feeding behavior (Strader et al. 2004); no phenotype in synovial endothelial cells (Patterson et al. 2005); enhanced long-term potentiation (LTP) in area CA1 (brain) and impaired performance in tasks assessing hippocampal function (Kaksonen et al. 2002); more sensitive to inhibition of food intake by the melanocortin agonist MTII (Reizes et al. 2003); perturbs laminar structure of the cerebral cortex as a result of impaired radial migration, and neural migration in the rostral migratory stream is impaired (Hienola et al. 2006); novel form of muscular dystrophy characterized by impaired locomotion, fibrosis, and hyperplasia of myonuclei and satellite cells (Cornelison et al. 2004).
Sdc4Syndecan-4Null allele: viable; enhanced fibrin deposition in degenerating fetal vessels in the placental labyrinth (Ishiguro et al. 2000); delayed angiogenesis in wound granulation tissue (Echtermeyer et al. 2001); defective subcellular localization of mTOR Complex2 and Akt activation in endothelial cells, affecting endothelial cell size, NOS, and arterial blood pressure (Partovian et al. 2008); decreased macrophage uptake of phospholipase A2-modified LDL (Boyanovsky et al. 2009); mesangial expansion, enhanced matrix collagens I and IV, fibronectin and focal segmental glomerulosclerosis in males, and induction of Sdc2 in glomeruli (Cevikbas et al. 2008); more susceptible to hepatic injury, and thrombin-cleaved form of osteopontin is significantly elevated after concanavalin-A injection (Kon et al. 2008); less damage in osteoarthritic cartilage in a surgically induced model of osteoarthritis (Echtermeyer et al. 2009); explanted satellite cells fail to reconstitute damaged muscle and are deficient in activation, proliferation, MyoD expression, myotube fusion, and differentiation (Cornelison et al. 2004); vibrissae are shorter and have a smaller diameter because of suboptimal response to fibroblast growth factors (Iwabuchi and Goetinck 2006); lower phosphorylation levels of focal adhesion kinase (Wilcox-Adelman et al. 2002); random migration of fibroblasts as a result of high delocalized Rac1 activity (Bass et al. 2007); defective RGD-independent cell attachment to transglutaminase-fibronectin matrices (Telci et al. 2008); impaired suppression of production of IL-1β by TGF-α (Ishiguro et al. 2002); decreased neutrophil recruitment and increased myofibroblast recruitment and interstitial fibrosis after bleomycin-treatment, no inhibition of fibrosis with recombinant CXCL10 protein (Jiang et al. 2010); hypersensitivity to LPS because of decreased TGF-β suppression of IL-1 production in monocytes and neutrophils (Ishiguro et al. 2001).
Gpc1Glypican-1Null allele: viable; reduced brain size (Jen et al. 2009). Athymic mutant mice show decreased tumor angiogenesis and metastasis (Aikawa et al. 2008).
Gpc2Glypican-2No mutants reported.
Gpc3Glypican-3Null allele: viable; resembles Simpson–Golabi–Behmel overgrowth syndrome, including somatic overgrowth, renal dysplasia, accessory spleens, polydactyly, and placentomegaly (Cano-Gauci et al. 1999; Chiao et al. 2002); defects in cardiac and coronary vascular development (Ng et al. 2009); alterations in Wnt signaling, in vivo inhibition of the noncanonical Wnt/JNK signaling, activation of canonical Wnt/β-catenin signaling (Song et al. 2005); increased Hedgehog signaling (Capurro et al. 2008); abnormal rates of proliferation and apoptosis in cortical and medullary collecting duct cells (Grisaru et al. 2001); delay in endochondral ossification, impairment in the development of the myelomonocytic lineage (Viviano et al. 2005).
Gpc4Glypican-4Zebrafish knypek controls cell polarity during convergent extension (Topczewski et al. 2001); craniofacial skeletal defects in adult fish (LeClair et al. 2009).
Gpc5Glypican-5No mutants reported.
Gpc6Glypican-6Impaired endochondral ossification and omodysplasia (Campos-Xavier et al. 2009).
Tgfbr3BetaglycanNull allele: embryonic lethal; heart and liver defects (Stenvers et al. 2003); defect in seminiferous cord formation in E12.5–13.5 embryos (Sarraj et al. 2010).
Hspg2PerlecanNull allele: embryonic lethal (E10–12); developmental angiogenesis altered in zebrafish (Zoeller et al. 2009); high incidence of malformations of the cardiac outflow tract, lack of well-defined spiral endocardial ridges (Costell et al. 2002); lower amounts of collagen IV and laminins in embryonic hearts, reduced function in infarcted hearts from heterozygous mice (Sasse et al. 2008); absence of acetylcholinesterase at the neuromuscular junctions (Arikawa-Hirasawa et al. 2002); cephalic and skeletal abnormalities (Arikawa-Hirasawa et al. 1999); cerebral ectopias, exencephaly (Girós et al. 2007); increased cross-sectional area of myosin heavy chain type IIb fibers in the tibialis anterior muscle (Xu et al. 2010b); diminished osteocyte canalicular pericellular area (Thompson et al. 2011).
Exon 3 deletion (Hspg23/3) viable: proteinuria after protein loading (Morita et al. 2005); monocyte/macrophage influx impaired in Hspg23/3Col18a1−/– mice in a model of renal ischemia/reperfusion (Celie et al. 2007).
Secreted as CSPG in some tissues (Danielson et al. 1992; Govindraj et al. 2002; Vogl-Willis and Edwards 2004; West et al. 2006), but relationship of CSPG isoform to phenotypes not established.
Prg1SerglycinNull allele: viable; secretory granule defects in mast cells (Abrink et al. 2004); dense core formation is defective in mast cell granules (Henningsson et al. 2006); defective secretory granule maturation and granzyme B storage in cytotoxic T cells (Grujic et al. 2005); no effect on macrophages (Zernichow et al. 2006); platelets and megakaryocytes contain unusual scroll-like membranous inclusions (Woulfe et al. 2008); enlargement of multiple lymphoid organs, decrease in the proportion of CD4+ cells, more pronounced airway inflammatory response in older mice (Wernersson et al. 2009); increased virulence of Klebsiella pneumoniae (Niemann et al. 2007); defective regulation of antiviral CD8+ T-cell responses (Grujic et al. 2008).
AgrnAgrinNull allele: embryonic lethal; reduced number, size, and density of postsynaptic acetylcholine receptor aggregates in muscles; abnormal intramuscular nerve branching and presynaptic differentiation (Gautam et al. 1996,1999); smaller brains (Serpinskaya et al. 1999); abnormal development of interneuronal synapses (Gingras et al. 2007); increased resistance to excitotoxic injury (Hilgenberg et al. 2002); reduced number of cortical presynaptic and postsynaptic specializations (Ksiazek et al. 2007).
Floxed allele: Inactivation in podocytes does not affect glomerular charge selectivity or glomerular architecture (Harvey et al. 2007).
Col18a1Collagen XVIIINull allele: viable; increased microvascular growth (Li and Olsen 2004); increased angiogenesis associated with atherosclerotic plaques (Moulton et al. 2004); delayed regression of blood vessels in the vitreous along the surface of the retina after birth and lack of or abnormal outgrowth of retinal vessels (Fukai et al. 2002); larger choroidal neovascularization lesions and increased vascular leakage (Marneros et al. 2007); accelerated healing and vascularization rate of excisional dorsal skin wounds (Seppinen et al. 2008); anomalous anastomoses of vasculature; disruption of the posterior iris pigment epithelial cell layer with release of melanin granules, severe thickening of the stromal iris basement membrane zone (Marneros and Olsen 2003); increase in the amount of retinal astrocytes (Hurskainen et al. 2005); more severe glomerular and tubulointerstitial injury in induced anti-GBM glomerulonephritis (Hamano et al. 2010); monocyte/macrophage influx impaired in Hspg23/3Col18a1−/– mice in a model of renal ischemia/reperfusion (Celie et al. 2007); mild chylomicronemia (Bishop et al. 2010).
Open in a separate window

Table 3.

Mouse mutants altered in HS biosynthesis
GeneEnzymePhenotype
Xt1Xylosyltransferase-1No mutants reported.
Xt2Xylosyltransferase-2Null allele: viable; polycystic kidney and livers (Condac et al. 2007).
GalTI (β4GalT7)Galactosyltransferase IHuman mutants: defective chondroitin substitution of decorin and biglycan in an Ehlers–Danlos patient (Gotte and Kresse 2005; Seidler et al. 2006).
GalTII (β3GalT6)Galactosyltransferase IINo mutants reported.
Glcat1Glucuronyltransferase INull allele: embryonic lethal (4–8-cell stage) (Izumikawa et al. 2010).
Extl3N-acetylglucosaminyl transferase IFloxed allele: Inactivation in islets decreases growth and insulin secretion (Takahashi et al. 2009).
Ext1/Ext2HS Copolymerase (N-acetylglucosaminyl-glucuronyltransferase)Null allele: embryonic lethal (E6-7.5); lack of mesoderm differentiation (Lin et al. 2000; Stickens et al. 2005); heterozygotes develop rib growth plate exostoses (Stickens et al. 2005; Zak et al. 2011); unaltered vascular permeability in heterozygous mice (Xu et al. 2010a).
Floxed allele of Ext1: defective brain morphogenesis and midline axon guidance after nestin-Cre inactivation (Inatani et al. 2003); no effect on adaptive immune response in CD15Cre mice (Garner et al. 2008); altered T-cell and dendritic cell homing to lymph nodes in Tie2Cre mice (Bao et al. 2010); rib growth plate exostosis formation in Col2Cre mice (Jones et al. 2010; Matsumoto et al. 2010; Zak et al. 2011).
Ndst1N-acetylglucosaminyl N-deacetylase/N-sulfotransferase-1Null allele: Perinatal lethal; lung hypoplasia, defective forebrain, lens, and skull development (Fan et al. 2000; Ringvall et al. 2000; Grobe et al. 2005; Pan et al. 2006).
Floxed allele: decreased chemokine transcytosis and presentation and neutrophil infiltration in Tie2Cre mice (Wang et al. 2005); decreased allergen-induced airway hyperresponsiveness and inflammation because of reduction in recruitment of eosinophils, macrophages, neutrophils, and lymphocytes in Tie2Cre mice (Zuberi et al. 2009); decreased pathological angiogenesis in Tie2Cre mice (Fuster et al. 2007); decreased vascular VEGF-induced hyperpermeability (Xu et al. 2010a); decreased vascular smooth muscle cell proliferation, vessel size, and vascular remodeling after arterial injury in SM22αCre mice (Adhikari et al. 2010a); mild effect on T-cell response in Tie2Cre;Ndst2−/−mice (Garner et al. 2008); defective lacrimal gland development and Fgf10-Fgfr2b complex formation and signaling in LeCre mice (Pan et al. 2008); defective lobuloalveolar development in mammary gland (Crawford et al. 2010).
Ndst2N-acetylglucosaminyl N-deacetylase/N-sulfotransferase-2Null allele: viable; mast cell deficiency and defective storage of proteases (Forsberg et al. 1999; Humphries et al. 1999); compounding mutation with Ndst1 reduces l-selectin interactions (Wang et al. 2005).
Ndst3N-acetylglucosaminyl N-deacetylase/N-sulfotransferase-3Null allele: viable; floxed allele available (Pallerla et al. 2008).
Ndst4N-acetylglucosaminyl N-deacetylase/N-sulfotransferase-4No mutants reported.
GlceUronyl C5 epimeraseNull allele: perinatal lethal; renal agenesis (Li et al. 2003).
H2stUronyl 2-O-sulfotransferaseNull allele: perinatal lethal; renal agenesis; skeletal and ocular defects (Bullock et al. 1998; Merry et al. 2001); defective cerebral cortical development (McLaughlin et al. 2003); altered lacrimal gland development (Qu et al. 2011).
Floxed allele: altered lipoprotein clearance in AlbCre mice (Stanford et al. 2010); altered branching morphogenesis in the mammary gland (Garner et al. 2011).
H3st1Glucosaminyl 3-O-sulfotransferase 1Null allele: partially penetrant lethality; no alteration in coagulation (HajMohammadi et al. 2003); fertility defects because of impaired ovarian function and placenta development (Shworak et al. 2002; HajMohammadi et al. 2003).
H3st2Glucosaminyl 3-O-sulfotransferase 2Null allele; viable; no neuronal phenotype (Hasegawa and Wang 2008).
H3st3Glucosaminyl 3-O-sulfotransferase 3No mutants reported.
H3st4Glucosaminyl 3-O-sulfotransferase 4No mutants reported.
H3st5Glucosaminyl 3-O-sulfotransferase 5No mutants reported.
H3st6Glucosaminyl 3-O-sulfotransferase 6No mutants reported.
H6st1Glucosaminyl 6-O-sulfotransferase 1Null allele: embryonic lethal (Habuchi et al. 2007; Sugaya et al. 2008).
Gene trap allele: embryonic lethal; retinal axon guidance defects (Pratt et al. 2006).
Floxed allele: systemic inactivation embryonic lethal (Izvolsky et al. 2008); no change in plasma triglycerides in AlbCre mice (Stanford et al. 2010).
H6st2Glucosaminyl 6-O-sulfotransferase 2Null allele: viable (Sugaya et al. 2008); HS6ST-2, but not HS6ST-1, morphants in zebrafish show abnormalities in the branching morphogenesis of the caudal vein (Chen et al. 2005).
H6st3Glucosaminyl 6-O-sulfotransferase 3No mutants reported.
HpaHeparanase, transgeneAccelerated wound angiogenesis, enhanced delayed type hypersensitivity response (Zcharia et al. 2005; Edovitsky et al. 2006; Ilan et al. 2006); accumulation of intracellular crystals of protein Ym1 in macrophages (Waern et al. 2010); resistance to amyloid protein A amyloidosis (Li et al. 2005); age-related enlargement of lymphoid tissue and altered leukocyte composition (Wernersson et al. 2009).
HpaHeparanaseNull allele: viable; altered MMP-2 and MMP-14 expression (Zcharia et al. 2009).
Sulf1Endo-6-sulfatase 1Null allele: viable; esophageal defect (Ai et al. 2007; Ratzka et al. 2008); enhanced osteoarthritis, MMP-13, ADAMTS-5, and noggin elevated, col2a1 and aggrecan reduced in cartilage and chondrocytes (Otsuki et al. 2010).
Sulf2Endo-6-sulfatase 2Null allele: viable; behavioral defects (Lamanna et al. 2006); enhanced osteoarthritis, MMP-13, ADAMTS-5, and noggin elevated, col2a1 and aggrecan reduced in cartilage and chondrocytes (Otsuki et al. 2010).
Gene trap allele: Sulf2GT(pGT1TMpfs)155Ska, no phenotype (Lum et al. 2007).
Open in a separate windowFigure 1 shows in pictorial form many of the systems in which HSPGs participate.
  1. HSPGs are present in basement membranes (perlecan, agrin, and collagen XVIII), where they collaborate with other matrix components to define basement membrane structure and to provide a matrix for cell migration.
  2. HSPGs are found in secretory vesicles, most notably serglycin, which plays a role in packaging granular contents, maintaining proteases in an active state, and regulating various biological activities after secretion such as coagulation, host defense, and wound repair.
  3. HSPGs can bind cytokines, chemokines, growth factors, and morphogens, protecting them against proteolysis. These interactions provide a depot of regulatory factors that can be liberated by selective degradation of the HS chains. They also facilitate the formation of morphogen gradients essential for cell specification during development and chemokine gradients involved in leukocyte recruitment and homing.
  4. HSPGs can act as receptors for proteases and protease inhibitors regulating their spatial distribution and activity.
  5. Membrane proteoglycans cooperate with integrins and other cell adhesion receptors to facilitate cell-ECM attachment, cell–cell interactions, and cell motility.
  6. Membrane HSPGs act as coreceptors for various tyrosine kinase-type growth factor receptors, lowering their activation threshold or changing the duration of signaling reactions.
  7. Membrane HSPGs act as endocytic receptors for clearance of bound ligands, which is especially relevant in lipoprotein metabolism in the liver and perhaps in the formation of morphogen gradients during development.
Open in a separate windowFigure 1.HSPGs have multiple activities in cells and tissues. (Adapted from Bishop et al. 2007; reprinted with permission from Nature Publishing Group © 2007.)This article is divided into 10 subsections. The first three are written for investigators outside the field who may need some background information on the diversity of HSPGs and the interactions that occur with protein ligands. The subsequent sections describe seven systems that illustrate general principles or ideas that have undergone a significant shift over the last decade. Because of space limitations not all subjects can be considered or treated in appropriate depth and therefore the reader is referred to excellent recent review articles (Tkachenko et al. 2005; Bulow and Hobert 2006; Bishop et al. 2007; Lamanna et al. 2007; Bix and Iozzo 2008; Filmus et al. 2008; Ori et al. 2008; Rodgers et al. 2008; Sanderson and Yang 2008; Iozzo et al. 2009; Couchman 2010).  相似文献   

12.
Membrane Domains Based on Ankyrin and Spectrin Associated with Cell–Cell Interactions     
Vann Bennett  Jane Healy 《Cold Spring Harbor perspectives in biology》2009,1(6)
Nodes of Ranvier and axon initial segments of myelinated nerves, sites of cell–cell contact in early embryos and epithelial cells, and neuromuscular junctions of skeletal muscle all perform physiological functions that depend on clustering of functionally related but structurally diverse ion transporters and cell adhesion molecules within microdomains of the plasma membrane. These specialized cell surface domains appeared at different times in metazoan evolution, involve a variety of cell types, and are populated by distinct membrane-spanning proteins. Nevertheless, recent work has shown that these domains all share on their cytoplasmic surfaces a membrane skeleton comprised of members of the ankyrin and spectrin families. This review will summarize basic features of ankyrins and spectrins, and will discuss emerging evidence that these proteins are key players in a conserved mechanism responsible for assembly and maintenance of physiologically important domains on the surfaces of diverse cells.Spectrins are flexible rods 0.2 microns in length with actin-binding sites at each end (Shotton et al. 1979; Bennett et al. 1982) (Fig. 1A). Spectrins are assembled from α and β subunits, each comprised primarily of multiple copies of a 106-amino acid repeat (Speicher and Marchesi 1984). In addition to the canonical 106-residue repeat, β spectrins also have a carboxy-terminal pleckstrin homology domain (Zhang et al. 1995; Macias et al. 1994) and tandem amino-terminal calponin homology domains (Bañuelos et al. 1998), whereas α spectrins contain an Src homology domain 3 (SH3) site (Musacchio et al. 1992), a calmodulin-binding site (Simonovic et al. 2006), and EF hands (Travé et al. 1995) (Fig. 1A). Spectrin α and β subunits are assembled antiparallel and side-to-side into heterodimers, which in turn are associated head-to-head to form tetramers (Clarke 1971; Shotton et al. 1979; Davis and Bennett 1983) (Fig. 1A). In human erythrocytes, in which spectrin was first characterized (Marchesi and Steers 1968; Clarke 1971), actin oligomers containing 10–14 monomers are each linked to five to six spectrin tetramers by accessory proteins to form a geodesic domelike structure that has been resolved by electron microscopy (Byers and Branton 1985). The principal proteins at the spectrin–actin junction are protein 4.1, adducin, tropomyosin, tropomodulin, and dematin (Bennett and Baines 2001) (Open in a separate windowFigure 1.Domain structure and variants of spectrin and ankyrin proteins. (A) Molecular domains of spectrins: Two α spectrins and five β spectrins are shown. Spectrins are comprised of modular units called spectrin repeats (yellow). Other domains such as the ankyrin binding domain (purple), Src-homology domain 3 (SH3, blue), EF-hand domain (red), and calmodulin-binding domain (green) promote interactions with binding targets important for spectrin function. The pleckstrin homology domain (black) promotes association with the plasma membrane and the actin binding domain (grey) tethers the spectrin-based membrane skeleton to the underlying actin cytoskeleton. (B) The spectrin tetramer, the fundamental unit of the spectrin-based membrane skeleton. The spectrin repeat domains of α and β spectrin associate end-to-end to form heterodimers. Heterodimers associate laterally in an antiparallel fashion to form tetramers. The tetramers can then associate end-to-end to form extended macromolecules that link into a geodesic dome shape directly underneath the plasma membrane. (C) Molecular domains present in canonical ankyrins. The membrane binding domain of ankyrin isoforms (orange) is comprised of 24 ANK repeats. The spectrin binding domain (green-blue) allows ankyrins to coordinate integral membrane proteins to the membrane skeleton. The death domain (pink) is the most highly conserved domain. The regulatory domain (brown) is the most variable region of ankyrins. The regulatory domain interacts intramolecularly with the membrane binding domain to modulate ankyrin’s affinity for other binding partners. All ankyrins and spectrins are subject to alternative splicing, which further increases their functional diversity.

Table 1.

Binding partners of spectrin and ankyrins
Spectrin Binding Partners
AlphaBeta
Transporters/ion channels
EnNaC (sodium)
NHE2 (ammonium)
Membrane anchors
PI lipids
Band 4.1
Ankyrin
EAAT4 (glutamate)
Membrane receptors
NMDA receptor
Signaling
RACK-1
Signaling
HsSH3pb1
Calmodulin
Cytoskeleton/cellular transport
F-actin
Adducin
Dynactin
Ankyrin Binding Partners
Membrane BDSpectrin BDDDREG D
Ion channels:
Anion exchanger
Na+/K+ATPase
Voltage-gated
Na+ channels
Na+/Ca2+ Exchanger
KCNG2/3
Rh antigen
IP3 receptor
Ryanodine receptor
Cell adhesion molecules:
L1-CAMs
CD44
E-cadherin
Dystroglycan
Cellular transport:
Tubulin
Clathrin
SpectrinFasLHsp40
Obscurin
PP2A
Open in a separate windowSpectrin is coupled to the inner surface of the erythrocyte membrane primarily through association with ankyrin, which is in turn linked to the cytoplasmic domains of the anion exchanger (Bennett 1978; Bennett and Stenbuck 1979a,b) and Rh/RhAG ammonium transporter (Nicolas et al. 2003). The spectrin-based membrane skeleton and its connections through ankyrin to membrane-spanning proteins are essential for survival of erythrocytes in the circulation, and mutations in these proteins result in hereditary hemolytic anemia (Bennett and Healy 2008). The ankyrin-binding sites of β spectrins 1–4 are located in the 15th spectrin repeat, which is folded identically to other repeats but has distinct surface-exposed residues (Davis et al. 2008; Ipsaro et al. 2009; Stabach et al. 2009) (Figs. 1A, A,2A).2A). Mammalian β-5 spectrin and its ortholog β-H spectrin in Drosophila and Caenorhabditis elegans are the only β spectrins lacking ankyrin-binding activity (Dubreuil et al. 1990; Thomas et al. 1998; McKeown et al. 1998; Stabach and Morrow 2000).Open in a separate windowFigure 2.Ankyrins and spectrins organize macromolecular complexes in diverse types of specialized membranes. (A) Ankyrin-G forms a complex with β-IV spectrin, neurofascin (a cell adhesion protein), and ion channels (KCNQ2/3 and voltage-gated sodium channel) at axon initial segments in Purkinje neurons. (B) In force buffering costameres of skeletal muscle, ankyrins -B and -G cooperate to target and stabilize key components of the dystroglycoprotein complex. At the membrane, ankyrin-G binds to dystrophin and β-dystroglycan. (C) In cardiomyocyte transverse tubules, ankyrins -B and -G coordinate separate microdomains. Ankyrin-B binds Na+/K+ ATPase, Na+/Ca2+ exchanger (NCX-1), and the inositol triphosphate receptor (IP3R). Ankyrin-G forms a complex with Nav1.5 and spectrin. (D) Ankyrin-G in epithelial lateral membrane assembly. Ankyrin-G binds to E-cadherin, β-2 spectrin, and the Na+/K+ ATPase. Spectrins are connected via F-actin bridges bound to α/γ adducin and tropomodulin.Ankyrin interacts with β spectrins through a ZU5 domain (Mohler et al. 2004a; Kizhatil et al. 2007a; Ipsaro et al. 2009) (Fig. 1B), and with most of its membrane partners through ANK repeats (Bennett and Baines 2001) (Fig. 2C,D). In addition, ankyrins have a highly conserve “death domain” and a carboxy-terminal regulatory domain (see the following discussion). The 24 ANK repeats are stacked in a superhelical array to form a solenoid (Michaely et al. 2002). Interestingly, the ANK repeat stack behaves like a reversible spring when stretched by atomic force microscopy, and may function in mechano-coupling in tissues such as the heart (Lee et al. 2006). ANK repeats are components of many proteins and participate in highly diverse protein interactions (Mosavi et al. 2004) (Fig. 2C). This versatile motif currently is being exploited using designed ANK repeat proteins (DARPins) engineered to interact with specific ligands that can function as substitutes for antibodies (Stumpp and Amstutz 2007; Steiner et al. 2008).Spectrin and ankyrin family members are expressed in most, if not all, animal (metazoan) cells, but are not present in bacteria, plants, or fungi. Spectrins are believed to have evolved from an ancestral α-actinin containing calponin homology domains and two spectrin repeats but not other domains (Thomas et al. 1997; Pascual et al. 1997). Ankyrin repeats are expressed in all phyla, presumably because of a combination of evolutionary relationships and in cases of bacteria and viruses by horizontal gene transfer. However, the spectrin-binding domain of ankyrin is present only in metazoans (Fig 1B). It is possible that evolution of ankyrins and spectrins could have been one of the adaptations required for organization of cells into tissues in multicellular animals.The human spectrin family includes two α subunits and five β subunits, whereas Drosophila and C. elegans have a single α subunit and two β subunits (Bennett and Baines 2001). Vertebrate ankyrins are encoded by three genes: ankyrin-R (ANK1) (the isoform first characterized in erythrocytes and also present in a restricted distribution in brain and muscle), ankyrin-B (ANK2), and ankyrin-G (ANK3). Vertebrate ankyrins evolved from a single gene in early chordates (Cai and Zhang 2006). C. elegans ankyrin is encoded by a single gene termed unc-44 (Otsuka et al. 1995), whereas the Drosophila genome contains two ankyrin genes: ankyrin (Dubreuil and Yu 1994) and ankyrin2 (Bouley et al. 2000).Mammalian ankyrins -B and -G are co-expressed in most cells, although they have distinct functions (Mohler et al. 2002; Abdi et al. 2006). Ankyrins -B and -G are closely related in their ANK repeats, and spectrin-binding domains, but diverge in their carboxy-terminal regulatory domains. Regulatory domains are natively unstructured and extended (Abdi et al. 2006). These flexible domains engage in intramolecular interactions with the membrane-binding and spectrin-binding domains (Hall and Bennett 1987; Davis et al. 1992; Abdi et al. 2006) that modulate protein associations and provide functional diversity between otherwise conserved ankyrins.In addition to the standard versions of ankyrins and spectrin subunits depicted in Figure 1, many variants of these proteins are expressed with the addition and/or deletion of functional domains because of alternative splicing of pre-mRNAs. For example, β spectrins can lack PH domains (Hayes et al. 2000), and giant ankyrins have insertions of up to 2000 residues (Kordeli et al. 1995; Chan et al. 1993; Pielage et al. 2008; Koch et al. 2008), whereas other ankyrins lack either the entire membrane-binding domain (Hoock et al. 1997), or both membrane- and spectrin-binding domains (Zhou et al. 1997). The insertions in 440 kDa ankyrin-B and 480 kDa ankyrin-G (Fig. 1B) have an extended conformation that potentially could have specialized roles in connections between the plasma membrane and cytoskeleton of axons where these giant ankyrins reside (Chan et al. 1993; Kordeli et al. 1995) (Fig. 1B). Interestingly, the inserted sequences in Drosophila giant ankyrins interact with microtubules at the presynaptic neuromuscular junction (Pielage et al. 2008) (see the following section).  相似文献   

13.
Mechanisms and Evidence of Genital Coevolution: The Roles of Natural Selection,Mate Choice,and Sexual Conflict     
Patricia L.R. Brennan  Richard O. Prum 《Cold Spring Harbor perspectives in biology》2015,7(7)
Genital coevolution between the sexes is expected to be common because of the direct interaction between male and female genitalia during copulation. Here we review the diverse mechanisms of genital coevolution that include natural selection, female mate choice, male–male competition, and how their interactions generate sexual conflict that can lead to sexually antagonistic coevolution. Natural selection on genital morphology will result in size coevolution to allow for copulation to be mechanically possible, even as other features of genitalia may reflect the action of other mechanisms of selection. Genital coevolution is explicitly predicted by at least three mechanisms of genital evolution: lock and key to prevent hybridization, female choice, and sexual conflict. Although some good examples exist in support of each of these mechanisms, more data on quantitative female genital variation and studies of functional morphology during copulation are needed to understand more general patterns. A combination of different approaches is required to continue to advance our understanding of genital coevolution. Knowledge of the ecology and behavior of the studied species combined with functional morphology, quantitative morphological tools, experimental manipulation, and experimental evolution have been provided in the best-studied species, all of which are invertebrates. Therefore, attention to vertebrates in any of these areas is badly needed.Of all the evolutionary interactions between the sexes, the mechanical interaction of genitalia during copulation in species with internal fertilization is perhaps the most direct. For this reason alone, coevolution between genital morphologies of males and females is expected. Morphological and genetic components of male and female genitalia have been shown to covary in many taxa (Sota and Kubota 1998; Ilango and Lane 2000; Arnqvist and Rowe 2002; Brennan et al. 2007; Rönn et al. 2007; Kuntner et al. 2009; Tatarnic and Cassis 2010; Cayetano et al. 2011; Evans et al. 2011, 2013; Simmons and García-González 2011; Yassin and Orgogozo 2013; and see examples in TaxaMale structuresFemale structuresEvidenceLikely mechanismReferencesMollusks Land snails (Xerocrassa)Spermatophore-producing organsSpermatophore-receiving organsComparative among speciesSAC or female choiceSauder and Hausdorf 2009 SatsumaPenis lengthVagina lengthCharacter displacementLock and keyKameda et al. 2009Arthropods Arachnids (Nephilid spiders)MultipleMultipleComparative among speciesSACKuntner et al. 2009 Pholcidae spidersCheliceral apophysisEpigynal pocketsComparative (no phylogenetic analysis)Female choiceHuber 1999 Harvestmen (Opiliones)Hardened penes and loss of nuptial giftsSclerotized pregenital barriersComparative among speciesSACBurns et al. 2013Millipedes Parafontaria tonomineaGonopod sizeGenital segment sizeComparative in species complexMechanical incompatibility resulting from Intersexual selectionSota and Tanabe 2010 Antichiropus variabilisGonopod shape and sizeAccesory lobe of the vulva and distal projectionFunctional copulatory morphologyLock and keyWojcieszek and Simmons 2012Crustacean Fiddler crabs, UcaGonopodeVulva, vagina, and spermathecaTwo-species comparison, shape correspondenceNatural selection against fluid loss, lock and key, and sexual selectionLautenschlager et al. 2010Hexapodes OdonatesClasping appendagesAbdominal shape and sensory hairsFunctional morphology, comparative among speciesLock and key via female sensory systemRobertson and Paterson 1982; McPeek et al. 2009Insects Coleoptera: seed beetlesSpiny aedagusThickened walls of copulatory ductComparative among speciesSACRönn et al. 2007 Callosobruchus: Callosobruchus maculatusDamage inflictedSusceptibility to damageFull sib/half sib mating experimentsSACGay et al. 2011Reduced spinesNo correlated responseExperimental evolutionSACCayetano et al. 2011 Carabid beetles (Ohomopterus)Apophysis of the endophallusVaginal appendix (pocket attached to the vaginal apophysis)Cross-species matingsLock and keySota and Kubota 1998; Sasabi et al. 2010 Dung beetle: Onthophagus taurusShape of the parameres in the aedagusSize and location of genital pitsExperimental evolutionFemale choiceSimmons and García-González 2011 Diptera: Drosophila santomea and D. yakubaSclerotized spikes on the aedagusCavities with sclerotized plateletsCross-species matingsSACKamimura 2012 Drosophila melanogaster species complexEpandrial posterior lobes
Oviscapt pouchesComparative among speciesSAC or female choiceYassin and Orgogozo 2013Phallic spikesOviscapt furrowsCercal teeth, phallic hook, and spinesUterine, vulval, and vaginal shields D. mauritiana and D. secheliaPosterior lobe of the genital archWounding of the female abdomenMating with introgressed linesSACMasly and Kamimura 2014 Stalk-eyed flies (Diopsidae)Genital processCommon spermathecal ductComparative among species and morphologicalFemale choiceKotrba et al. 2014 Tse-tse flies: Glossina pallidipesCercal teethFemale-sensing structuresExperimental copulatory functionFemale choiceBriceño and Eberhard 2009a,b Phelebotomine: sand fliesAedagal filaments, aedagal sheathsSpermathecal ducts length, base of the ductComparative among speciesNone specifiedIlango and Lane 2000 Heteroptera: Bed bugs (Cimiciidae)Piercing genitaliaSpermalege (thickened exosqueleton)Comparative among speciesSACCarayon 1966; Morrow and Arnqvist 2003 Plant bugs (Coridromius)Changes in male genital shapeExternal female paragenitaliaComparative among speciesSACTatarnic and Cassis 2010 Waterstriders (Gerris sp.)Grasping appendagesAntigrasping appendagesComparative among speciesSACArnqvist and Rowe 2002 Gerris incognitusGrasping appendagesAntigrasping appendagesComparative among populationsSACPerry and Rowe 2012 Bee assassins (Apiomerus)AedagusBursa copulatrixComparative among speciesNoneForero et al. 2013 Cave insects (Psocodea), NeotroglaMale genital chamberPenis-like gynosomeComparative among speciesFemale competition (role reversal), coevolution SACYoshizawa et al. 2014 Butterflies (Heliconiinae)Thickness of spermatophore wallSigna: Sclerotized structure to break spermatophoresComparative among speciesSACSánchez and Cordero 2014Fish Basking shark: Cetorhinus maximusClasper clawThick vaginal padsMorphological observationNoneMatthews 1950 GambusiaGonopodial tipsGenital papillae within openingsComparative among speciesStrong character displacementLangerhans 2011 Poecilia reticulataGonopodium tip shapeFemale gonopore shapeComparative among populationsSACEvans et al. 2011Reptiles AnolesHemipene shapeVagina shapeShape correspondence, two speciesSexual selectionKöhler et al. 2012 Several speciesHemipene shapeVagina shapeShape correspondenceLock and key, female choice, and SACPope 1941; Böhme and Ziegler 2009; King et al. 2009 Asiatic pit vipersSpininess in hemipenesThickness of vagina wallTwo-species comparisonNonePope 1941 Garter snake: Thamnophis sirtalisBasal hempene spineVaginal muscular controlExperimental manipulationSACFriesen et al. 2014Birds WaterfowlPenis lengthVaginal elaborationComparative among speciesSACBrennan et al. 2007 TinamousPenis length/presenceVaginal elaborationComparative among speciesFemale choice/natural selectionPLR Brennan, K Zyscowski, and RO Prum, unpubl.Mammals MarsupialsBifid penisTwo lateral vaginaeShape correspondenceNoneRenfree 1987 EquidnaBifid penis with four rosettesSingle vagina splits into two uteriShape correspondenceNoneAugee et al. 2006; Johnston et al. 2007 Insectivores: Short-tailed shrew: Blarina brevicaudaS-shaped curve of the erect penisCoincident curve in the vaginaShape correspondenceNoneBedford et al. 2004 Common tenrec: Tenrec caudatusFiliform penis (up to 70% of the male’s body length)Internal circular folds in the vaginaLength correspondenceNoneBedford et al. 2004 Rodents: Cape dune mole: Bathyergus suillusPenis and baculum lengthVaginal lengthAllometric relationships within speciesNoneKinahan et al. 2007 Australian hopping mice (Notomys)Spiny penisDerived distal region in the vaginaMorphological observation and two-species comparisonCopulatory lockBreed et al. 2013 Pig: Sus domesticusFiliform penis endCervical ridgesArtificial inseminationFemale choiceBonet et al. 2013 Primates: Macaca arctoidesLong and filamentous glansVestibular colliculus (fleshy fold) that partially obstructs the entrance to the vaginaShape correspondence and comparison with close relativesNoneFooden 1967
Open in a separate windowThe likely mechanism is that suggested by the authors, and it includes sexually antagonistic coevolution (SAC), natural selection, sexual selection, female choice, or none specified. The evidence provided by the studies can be comparative among species or among populations, experimental evolution, cross-species matings, full-sibling (sib)/half-sib matings, shape, and length correspondence. Shape correspondence is often taken as evidence of coevolution, although it is not as conclusive as other approaches.Male genitalia are among the most variable structures in nature (Eberhard 1985). In contrast, female genitalia have typically been found not to be as interspecifically variable as male genitalia in several studies that specifically examined and described them (Eberhard 1985, 2010a,b). Female genitalia are not studied as often as male genitalia, perhaps because of a male-biased view of evolutionary processes by researchers (Ah-King et al. 2014). However, studying female genitalia is undeniably challenging. Male genitalia are generally kept inside of the body cavity, but are everted before, or during copulation, so their functional morphology can be more easily studied than the internal genitalia of females. Female genitalia also tend to be softer than male genitalia and thus their morphology may be more difficult to describe, and can more easily be distorted on dissection and preservation. Female adaptations to sense or oppose features of male genitalia can be subtle, requiring careful study. Female genital tracts are under multiple sources of selection: not just mating, but also storing sperm, egg laying, birthing, and often interfacing with the terminal portion of the digestive tract. Therefore, selection balancing multiple functions may further constrain morphological evolution in female genitalia. However, even small morphological changes in female genitalia, for example, increases in vaginal muscle, may change a female’s ability to choose or reject a male during mating, or to manage the costs of mating. Thus, the functional consequences to male and female genital morphology are hard to predict unless one knows how genitalia function during intromission. Despite these challenges, recent studies have examined variation of female genitalia and evidence is accumulating that features of female genitalia are variable enough to support coevolutionary processes (Polihronakis 2006; Puniamoorthy et al. 2010; Siegel et al. 2011; Showalter et al. 2013; and see additional references in Ah-King et al. 2014).In this article, we will discuss different hypotheses of genital evolution that predict coevolution; however, this is not a review of that entire subject (but see Eberhard et al. 2010b; Simmons 2013). Rather, we discuss the various mechanisms of genital coevolution differentiating the potentially independent or overlapping roles of natural selection, female choice, and male–male competition (Fig. 1). This classification allows us to distinguish specifically those mechanisms of genital coevolution that involve sexual conflict (i.e., when the evolutionary interests of individuals of different sexes, particularly over mating, are different). We then highlight examples in different taxa organisms with particular emphasis on those that provide evidence of sexual conflict.Open in a separate windowFigure 1.Graphical classification of mechanisms of genital evolution and coevolution. Three circles depict the independent and co-occurring actions of natural selection, female choice, and male–male competition. Different specific versions of genital coevolution can occur depending on which of the three broader evolutionary mechanisms are occurring. Sexual conflict (hatched lines) occurs through the simultaneous action of male–male competition and female choice, or male–male competition and natural selection. SAC, sexually antagonistic coevolution. See text for explanation.  相似文献   

14.
Wnt/Wingless Signaling in Drosophila     
Sharan Swarup  Esther M. Verheyen 《Cold Spring Harbor perspectives in biology》2012,4(6)
  相似文献   

15.
Natural Killer Cell Tolerance: Control by Self or Self-Control?     
Baptiste N. Jaeger  Eric Vivier 《Cold Spring Harbor perspectives in biology》2012,4(3)
  相似文献   

16.
Signaling by Nuclear Receptors     
Richard Sever  Christopher K. Glass 《Cold Spring Harbor perspectives in biology》2013,5(3)
  相似文献   

17.
A Systematic Proteomic Analysis of Listeria monocytogenes House-keeping Protein Secretion Systems     
Sven Halbedel  Swantje Reiss  Birgit Hahn  Dirk Albrecht  Gopala Krishna Mannala  Trinad Chakraborty  Torsten Hain  Susanne Engelmann  Antje Flieger 《Molecular & cellular proteomics : MCP》2014,13(11):3063-3081
  相似文献   

18.
Helicase Loading at Chromosomal Origins of Replication     
Stephen P. Bell  Jon M. Kaguni 《Cold Spring Harbor perspectives in biology》2013,5(6)
Loading of the replicative DNA helicase at origins of replication is of central importance in DNA replication. As the first of the replication fork proteins assemble at chromosomal origins of replication, the loaded helicase is required for the recruitment of the rest of the replication machinery. In this work, we review the current knowledge of helicase loading at Escherichia coli and eukaryotic origins of replication. In each case, this process requires both an origin recognition protein as well as one or more additional proteins. Comparison of these events shows intriguing similarities that suggest a similar underlying mechanism, as well as critical differences that likely reflect the distinct processes that regulate helicase loading in bacterial and eukaryotic cells.Replicative DNA helicases are ring-shaped molecules with a central cavity through which DNA passes as they unwind DNA. Their loading at replication origins is a critical and highly regulated event in chromosomal replication. The DNA helicase is the first of the replication fork proteins recruited to and loaded onto origins of replication, and the loaded helicase is required for the recruitment of the rest of the replication machinery (Remus and Diffley 2009; Kaguni 2011). Indeed, the replicative DNA helicase links the replication machinery to the parental DNA (O’Donnell 2006). In Escherichia coli cells, the DnaB replicative helicase binds to primase (DnaG) and the sliding clamp loader, which in turn binds the DNA polymerases. Although the polymerases are also linked to the template DNA by sliding clamps, when these interactions are broken, the polymerases’ association with the sliding clamp loader and the helicase keeps them at the site of replication. The interactions that tether the DNA polymerases to the eukaryotic replication fork are less clear but very likely involve direct and indirect interactions with the Mcm2–7 replicative helicase (Calzada et al. 2005).Helicase loading is carefully regulated to control the location and frequency of replication initiation. In eukaryotic cells, helicase loading is tightly restricted to the G1 phase of the cell cycle. This constraint is a key part of the mechanisms that ensure that no origin can initiate more than once per cell cycle (Siddiqui et al. 2013; Zielke et al. 2013). In addition, the sites of eukaryotic replicative helicase loading define the potential sites of replication initiation in the cell (but not all loaded helicases are used during a given S phase [Rhind and Gilbert 2012]). Although the central regulated event in bacterial chromosome duplication is the recruitment of the ATP-bound initiator protein DnaA (Skarstad and Katayama 2013), the loading of the replicative helicase represents a key committed step during initiation.Here we will discuss the mechanism of helicase loading in bacteria and eukaryotic cells. Much of the discussion will focus on studies in the bacterium E. coli, the yeast Saccharomyces cerevisiae, and the frog Xenopus laevis, in which the events of helicase loading are best understood. Comparison of these mechanisms shows important similarities and differences between the domains of life. In both bacteria and eukaryotic cells, multiple AAA+ proteins use ATP binding and hydrolysis to direct helicase loading and both helicases are initially loaded in an inactive form. On the other hand, the eukaryotic helicase is loaded around double-stranded DNA (dsDNA) and as a double hexamer, whereas the bacterial helicase is loaded around single-stranded DNA (ssDNA) as a single hexamer. These distinctions are very likely due to the very different regulatory mechanisms of DNA replication in bacteria and eukaryotic cells.Before describing the process of helicase loading, it is relevant to know that the essential function of replicative helicases is to unwind the parental duplex DNA using the energy provided by the hydrolysis of a nucleoside triphosphate (Patel et al. 2011). Replicative DNA polymerases then copy each parental DNA strand to duplicate the genome. That replicative DNA helicases are hexameric (and sometimes heptameric) structures that have a positively charged, central channel provides a framework for how these molecular machines work. Several models that describe the mechanism of unwinding have been considered. One is the steric or strand-exclusion model, in which the helicase excludes one strand of DNA while the other passes through the central cavity as the enzyme moves. Because these enzymes can bind and translocate on duplex DNA without unwinding (Kaplan et al. 2003), two additional models have been proposed. In the ploughshare model, duplex DNA enters the central cavity and exits as unwound DNA by virtue of a domain/protein that acts as a ploughshare or pin, which disrupts the hydrogen bonds of DNA as it is pumped through the enzyme (reviewed in Takahashi et al. 2005). A second model, the DNA-pumping model, was proposed on the basis of the double-hexameric forms of SV40, Mcm2–7, and other replicative helicases (Mastrangelo et al. 1989; Remus et al. 2009). This model proposes that the two helicases pump duplex DNA toward one another, resulting in torsional strain that forces the two strands apart, at which point they exit the central channel as two ssDNA loops. Current evidence supports the steric exclusion model (Jezewska et al. 1998b; Kaplan 2000; Galletto et al. 2004; Fu et al. 2011). Of interest, the E. coli enzyme moves in the 5′-to-3′ direction relative to the engaged ssDNA that passes through the central cavity, whereas archaeal and eukaryotic enzymes move in the 3′-to-5′ direction.

Table 1.

Replicative DNA helicases of free-living organisms are hexameric
DomainModel organism(s)HelicaseaDirection of movement
BacteriaEscherichia coliDnaB5′ → 3′
ArchaeaSulfolobus solfataricusMCM3′ → 5′
EukaryaSaccharomyces cerevisiae
Drosophila melanogaster
Xenopus laevis
Mcm2–73′ → 5′
Open in a separate windowaDnaB and S. solfataricus MCM are homohexamers, whereas eukaryotic Mcm2–7 is composed of six nonidentical subunits.  相似文献   

19.
How Do Astrocytes Participate in Neural Plasticity?     
Philip G. Haydon  Maiken Nedergaard 《Cold Spring Harbor perspectives in biology》2015,7(3)
Work over the past 20 years has implicated electrically nonexcitable astrocytes in complex neural functions. Despite controversies, it is increasingly clear that many, if not all, neural processes involve astrocytes. This review critically examines past work to identify the commonalities among the many published studies of neuroglia signaling. Although several studies have shown that astrocytes can impact short-term and long-term synaptic plasticity, further work is required to determine the requirement for astrocytic Ca2+ and other second messengers in these processes. One of the roadblocks to the field advancing at a rapid pace has been technical. We predict that the novel experimental tools that have emerged in recent years will accelerate the field and likely disclose an entirely novel path of neuroglia signaling within the near future.The year 2014 represents the 20th anniversary of a pair of papers published by the writers that provided the first indication that astrocytes actively signal to neurons, and that these glial cells have the potential to be participants in the control of neural circuit function and behavior (Nedergaard 1994; Parpura et al. 1994). Although we have taken independent paths, we come together on this anniversary to discuss what we have learned since 1994, where we see the field progressing, and, finally, to discuss some key steps that we believe should be taken in the next decade to begin to further clarify the diverse roles that astrocytes play in brain function in health and disease.Without knowledge of one another’s work, we published a pair of papers that provided the first demonstration that physiological changes in astrocytes influence neurons: stimulated Ca2+ changes in astrocytes led to delayed Ca2+ responses in neurons (Nedergaard 1994; Parpura et al. 1994). To set the backdrop to these studies, this was a period of explosive growth in Ca2+ imaging that resulted from the availability of fluorescent Ca2+ indicators and sensitive cameras to permit low-light-level imaging of intracellular biochemistry. As a consequence, there had been several recently published studies revealing that astrocytes show Ca2+ elevations in response to mechanical contact or even to the addition of neurotransmitters, such as glutamate (Cornell-Bell et al. 1990). However, the functional downstream consequences were unknown. In our studies, which used a combination of different stimuli—mechanical, optical, as well as a chemical transmitter bradykinin—we were able to show a robust impact of the astrocytic Ca2+ signal on the adjacent neurons (Nedergaard 1994; Parpura et al. 1994). Despite differences in mechanistic conclusions, these papers stimulated revised thinking about roles of astrocytes in the brain—perhaps these glial cells were actively signaling, albeit on a slower time scale, to modulate neurons, circuits, and, ultimately, behavior.The subsequent two decades have been spent examining the signaling of astrocytes to neurons in more intact systems. As a consequence of this work, it is clear that astrocytes play critical roles in actively modulating brain function, although we are still putting the pieces of the puzzle in place to understand where, when, and how this process occurs naturally in vivo and when dysfunction can lead to disorders of the brain.The field has gone through an explosive growth that has consisted of several phases. Initially, there was the dish phase in which the potential of the astrocyte was revealed in cell culture. These studies were essential as they captured our imagination and stimulated new thinking; however, they were limited by the fact that the properties of astrocytes can be different in vitro and in vivo. Next, we had the in situ phase, in which we asked whether similar processes could be detected in situ in acutely isolated brain slices. Then we began asking about roles in vivo through Ca2+ imaging together with two-photon microscopy as well as with molecular genetic alterations to permit the inhibition and stimulation of astrocytes. Each of these phases has offered unique insights, challenges, and opportunities for the field.Through all of these phases, an emergent picture is developing in which it is without doubt that astrocytes play important roles in vivo but that the mechanisms are so diverse and complex that an understanding is still to emerge.Some of the major challenges that we have faced and continue to face include: What are the endogenous signals of these glial cells? How can we stimulate astrocytes in a physiologically relevant manner? How can we inhibit astrocytes to determine when they are needed for brain function? And last, but by no means least, how diverse are astrocytes?We are still at the early days of understanding astrocytes, and patience regarding functional interpretation is required. We make this statement because there have been apparently contradictory conclusions drawn from different studies. The individual observations are important as they help provide a fuller picture of the biology of these complex cells. However, the jury still needs further evidence before definitive conclusions can be drawn. For example, a plethora of studies have shown that Ca2+ signals stimulate gliotransmission and the consequent modulation of neurons and synapses (see Agulhon et al. 2010), the field was rocked. We do not question the data of the study; instead, we believe this is an important piece of information that ultimately needs to be put in context. In contrast, additional studies have shown that IP3 receptors are important for other aspects of astrocyte-induced synaptic modulation. A more recent study has shown that, in addition to Ca2+ release from internal stores, the influx of Ca2+ through transient receptor potential (TRP) channels is important for gliotransmission (Shigetomi et al. 2013). Clearly, such influx sites were unlikely to have been affected by the IP3R2 knockout and were shown to regulate d-serine release from the astrocyte. Another piece of the puzzle is added. Undoubtedly, there will be further twists and turns, but that is the joy of discovery, and it should be embraced.

Table 1.

Effect of Ca2+ signaling on excitatory or inhibitory potentials, slow inward current, synaptic failure, or neural bistability
PreparationMethod for inducing astrocytic Ca2+ signalingChange in the frequency of EPSP or IPSP (%)aDuration of modulation of EPSP or IPSPaReferences
Hippocampal coculturesMechanical stimulation or photolysis of caged Ca2+-10–50 secAraque et al. 1998
RetinaMechanical stimulationModulation of light- induced neural activity10–20 secNewman and Zahs 1998
Frog neuromuscular junctionInjection of GTP-γS in perisynaptic Schwann cellsModulation of nerve-evoked synaptic responsesFor the duration of the recordingsRobitaille 1998
Hippocampal slicesTrain of depolarization10%–30%∼60–120 secJourdain et al. 2007
Hippocampal slicesPhotolysis20%–30% decrease in synaptic failure50–60 secPerea and Araque 2007
Hippocampal slicesAgonists (ATP, UTP, FMRF)∼20%–30%10–60 secWang et al. 2012a
Hippocampal slicesPhotolysis of caged Ca2+b∼30%10–60 secWang et al. 2013
Slow inward current
Hippocampal slicesAgonist (DHPG) and photolysis of caged Ca2+Slow inward current∼20–50 secFellin et al. 2004
Hippocampal slicesNeuronal depolarizationSlow inward current∼50 secNavarrete and Araque 2008
Decrease in synaptic failure rate
Hippocampal slicesTrain of depolarization∼20%–30%20 minKang et al. 1998
Hippocampal slicesPhotolysis of caged Ca2+∼20%–30%∼60 secPerea and Araque 2007
Hippocampal slicesAgonists∼20%–30%10–60 secWang et al. 2012a
Hippocampal slicesComparison of agonists and photolysis of caged Ca2+∼20%–30%10–60 secWang et al. 2013
Bistability
Cortical slicesTrains of depolarizationUpstate synchronizations-Poskanzer and Yuste 2011
Cerebellar slicesAgonists (ATP, UTP, FMRF)Increase in duration of upstate40–60 secWang et al. 2012b
Open in a separate windowAll of the studies included in the table show that the modulatory effect on neural activity is Ca2+-dependent (BAPTA loading, thapsigargin, and/or use of transgenic mice with deletion of IP3R2 receptors).DHPG, dihydroxyphenylglycine; EPSP, excitatory postsynaptic potential; IPSP, inhibitory postsynaptic potential.aFor simplicity, EPSP and IPSP denote excitatory or inhibitory potentials or currents in both presence and absence of tetrodotoxin (TTX). Details can be found in the original papers.bThe same study compared the effect of photolysis and agonist-induced astrocytic Ca2+ signaling and found that only photolysis, but not agonist exposure, induced changes in the frequency of EPSPs.We believe that it is also important to be constrained when discussing Ca2+ as there is not just one type of Ca2+ signal. For example, there are global Ca2+ signals in which large somatic Ca2+ elevations arise and that can propagate as slow waves between adjacent astrocytes in slices preparation. In vivo, astrocytes in awake mice display global Ca2+ increases that often simultaneously engaged most cells within the field of view. Isolated oscillatory cellular Ca2+ signals can be restricted to one cell, and there are “spotty” Ca2+ signals that can be restricted to local microdomains (Shigetomi et al. 2013). It is possible, even likely, that each of these signals mediates different processes and more effort should focus on understanding the important functional distinction between each.Other areas of interesting debate have concerned how gliotransmitters are released. Evidence exists for multiple mechanisms: exocytosis, anion transporters, and connexin hemichannels, to name a few. Significant evidence exists for each, and it is likely that all are used, although in different locales, and are recruited under differing conditions. A challenge is to perform precise experiments that allow the discrimination between each mechanism and to identify when each is recruited in physiology and/or pathology.  相似文献   

20.
Variation in Adult Plant Phenotypes and Partitioning among Seed and Stem-Borne Roots across Brachypodium distachyon Accessions to Exploit in Breeding Cereals for Well-Watered and Drought Environments     
Vincent Chochois  John P. Vogel  Gregory J. Rebetzke  Michelle Watt 《Plant physiology》2015,168(3):953-967
Seedling roots enable plant establishment. Their small phenotypes are measured routinely. Adult root systems are relevant to yield and efficiency, but phenotyping is challenging. Root length exceeds the volume of most pots. Field studies measure partial adult root systems through coring or use seedling roots as adult surrogates. Here, we phenotyped 79 diverse lines of the small grass model Brachypodium distachyon to adults in 50-cm-long tubes of soil with irrigation; a subset of 16 lines was droughted. Variation was large (total biomass, ×8; total root length [TRL], ×10; and root mass ratio, ×6), repeatable, and attributable to genetic factors (heritabilities ranged from approximately 50% for root growth to 82% for partitioning phenotypes). Lines were dissected into seed-borne tissues (stem and primary seminal axile roots) and stem-borne tissues (tillers and coleoptile and leaf node axile roots) plus branch roots. All lines developed one seminal root that varied, with branch roots, from 31% to 90% of TRL in the well-watered condition. With drought, 100% of TRL was seminal, regardless of line because nodal roots were almost always inhibited in drying topsoil. Irrigation stimulated nodal roots depending on genotype. Shoot size and tillers correlated positively with roots with irrigation, but partitioning depended on genotype and was plastic with drought. Adult root systems of B. distachyon have genetic variation to exploit to increase cereal yields through genes associated with partitioning among roots and their responsiveness to irrigation. Whole-plant phenotypes could enhance gain for droughted environments because root and shoot traits are coselected.Adult plant root systems are relevant to the size and efficiency of seed yield. They supply water and nutrients for the plant to acquire biomass, which is positively correlated to the harvest index (allocation to seed grain), and the stages of flowering and grain development. Modeling in wheat (Triticum aestivum) suggested that an extra 10 mm of water absorbed by such adult root systems during grain filling resulted in an increase of approximately 500 kg grain ha−1 (Manschadi et al., 2006). This was 25% above the average annual yield of wheat in rain-fed environments of Australia. This number was remarkably close to experimental data obtained in the field in Australia (Kirkegaard et al., 2007). Together, these modeling and field experiments have shown that adult root systems are critical for water absorption and grain yield in cereals, such as wheat, emphasizing the importance of characterizing adult root systems to identify phenotypes for productivity improvements.Most root phenotypes, however, have been described for seedling roots. Seedling roots are essential for plant establishment, and hence, the plant’s potential to set seed. For technical reasons, seedlings are more often screened than adult plants because of the ease of handling smaller plants and the high throughput. Seedling-stage phenotyping may also improve overall reproducibility of results because often, growth media are soil free. Seedling soil-free root phenotyping conditions are well suited to dissecting fine and sensitive mechanisms, such as lateral root initiation (Casimiro et al., 2003; Péret et al., 2009a, 2009b). A number of genes underlying root processes have been identified or characterized using seedlings, notably with the dicotyledonous models Arabidopsis (Arabidopsis thaliana; Mouchel et al., 2004; Fitz Gerald et al., 2006; Yokawa et al., 2013) and Medicago truncatula (Laffont et al., 2010) and the cereals maize (Zea mays; Hochholdinger et al., 2001) and rice (Oryza sativa; Inukai et al., 2005; Kitomi et al., 2008).Extrapolation from seedling to adult root systems presents major questions (Hochholdinger and Zimmermann, 2008; Chochois et al., 2012; Rich and Watt, 2013). Are phenotypes in seedling roots present in adult roots given developmental events associated with aging? Is expression of phenotypes correlated in seedling and adult roots if time compounds effects of growth rates and growth conditions on roots? Watt et al. (2013) showed in wheat seedlings that root traits in the laboratory and field correlated positively but that neither correlated with adult root traits in the field. Factors between seedling and adult roots seemed to be differences in developmental stage and the time that growing roots experience the environment.Seedling and adult root differences may be larger in grasses than dicotyledons. Grass root systems have two developmental components: seed-borne (seminal) roots, of which a number emerge at germination and continue to grow and branch throughout the plant life, and stem-borne (nodal or adventitious) roots, which emerge from around the three-leaf stage and continue to emerge, grow, and branch throughout the plant life. Phenotypes and traits of adult root systems of grasses, which include the major cereal crops wheat, rice, and maize, are difficult to predict in seedling screens and ideally identified from adult root systems first (Gamuyao et al., 2012).Phenotyping of adult roots is possible in the field using trenches (Maeght et al., 2013) or coring (Wasson et al., 2014). A portion of the root system is captured with these methods. Alternatively, entire adult root systems can be contained within pots dug into the ground before sowing. These need to be large; field wheat roots, for example, can reach depths greater than 1.5 m depending on genotype and environment. This method prevents root-root interactions that occur under normal field sowing of a plant canopy and is also a compromise.A solution to the problem of phenotyping adult cereal root systems is a model for monocotyledon grasses: Brachypodium distachyon. B. distachyon is a small-stature grass with a small genome that is fully sequenced (Vogel et al., 2010). It has molecular tools equivalent to those available in Arabidopsis (Draper et al., 2001; Brkljacic et al., 2011; Mur et al., 2011). The root system of B. distachyon reference line Bd21 is more similar to wheat than other model and crop grasses (Watt et al., 2009). It has a seed-borne primary seminal root (PSR) that emerges from the embryo at seed germination and multiple stem-borne coleoptile node axile roots (CNRs) and leaf node axile roots (LNRs), also known as crown roots or adventitious roots, that emerge at about three leaves through to grain development. Branch roots emerge from all root types. There are no known anatomical differences between root types of wheat and B. distachyon (Watt et al., 2009). In a recent study, we report postflowering root growth in B. distachyon line Bd21-3, showing that this model can be used to answer questions relevant to the adult root systems of grasses (Chochois et al., 2012).In this study, we used B. distachyon to identify adult plant phenotypes related to the partitioning among seed-borne and stem-borne shoots and roots for the genetic improvement of well-watered and droughted cereals (Fig. 1; Krassovsky, 1926; Navara et al., 1994), nitrogen, phosphorus (Tennant, 1976; Brady et al., 1995), oxygen (Wiengweera and Greenway, 2004), soil hardness (Acuna et al., 2007), and microorganisms (Sivasithamparam et al., 1978). Of note is the study by Krassovsky (1926), which was the first, to our knowledge, to show differences in function related to water. Krassovsky (1926) showed that seminal roots of wheat absorbed almost 2 times the water as nodal roots per unit dry weight but that nodal roots absorbed a more diluted nutrient solution than seminal roots. Krassovsky (1926) also showed by removing seminal or nodal roots as they emerged that “seminal roots serve the main stem, while nodal roots serve the tillers” (Krassovsky, 1926). Volkmar (1997) showed, more recently, in wheat that nodal and seminal roots may sense and respond to drought differently. In millet (Pennisetum glaucum) and sorghum (Sorghum bicolor), Rostamza et al. (2013) found that millet was able to grow nodal roots in a dryer soil than sorghum, possibly because of shoot and root vigor.Open in a separate windowFigure 1.B. distachyon plant scanned at the fourth leaf stage, with the root and shoot phenotypes studied indicated. Supplemental Table S1.
PhenotypeAbbreviationUnitRange of Variation
All Experiments (79 Lines and 582 Plants)Experiment 6 (36 Lines)
Whole plant
TDWTDWMilligrams88.6–773.8 (×8.7)285.6–438 (×1.5)
Shoot
SDWSDWMilligrams56.4–442.5 (×7.8)78.2–442.5 (×5.7)
 No. of tillersTillerNCount2.8–20.3 (×7.4)10–20.3 (×2)
Total root system
TRLTRLCentimeters1,050–10,770 (×10.3)2,090–5,140 (×2.5)
RDWRDWMilligrams28.9–312.17 (×10.8)62.2–179.1 (×2.9)
RootpcRootpcPercentage (of TDW)20.5–60.6 (×3)20.5–44.3 (×2.2)
R/SR/SUnitless ratio0.26–1.54 (×6)0.26–0.80 (×3.1)
PSRs
 Length (including branch roots)PSRLCentimeters549.1–4,024.6 (×7.3)716–2,984 (×4.2)
PSRpcPSRpcPercentage (of TRL)14.9–94.1 (×6.3)31.3–72.3 (×2.3)
 No. of axile rootsPSRcountCount11
 Length of axile rootPSRsumCentimeters17.45–52 (×3)17.45–30.3 (×1.7)
 Branch rootsPSRbranchCentimeters · (centimeters of axile root)−119.9–109.3 (×5.5)29.3–104.3 (×3.6)
CNRs
 Length (including branch roots)CNRLCentimeters0–3,856.70–2,266.5
CNRpcCNRpcPercentage (of TRL)0–57.10–49.8
 No. of axile rootsCNRcountCount0–20–2
 Cumulated length of axile rootsCNRsumCentimeters0–113.90–47.87
 Branch rootsCNRbranchCentimeters · (centimeters of axile root)−10–77.80–77.8
LNRs
 Length (including branch roots)LNRLCentimeters99.5–5,806.5 (×58.5)216.1–2,532.4 (×11.7)
LNRpcLNRpcPercentage (of TRL)4.2–72.7 (×17.5)6–64.8 (×10.9)
LNRcountLNRcountCount2–22.2 (×11.1)3.3–15.3 (×4.6)
LNRsumLNRsumCentimeters25.9–485.548–232 (×4.8)
 Branch rootsLNRbranchCentimeters · (centimeters of axile root)−12.1–25.4 (×12.1)3.2–15.9 (×5)
Open in a separate windowThe third reason for dissecting the different root types in this study was that they seem to have independent genetic regulation through major genes. Genes affecting specifically nodal root growth have been identified in maize (Hetz et al., 1996; Hochholdinger and Feix, 1998) and rice (Inukai et al., 2001, 2005; Liu et al., 2005, 2009; Zhao et al., 2009; Coudert et al., 2010; Gamuyao et al., 2012). Here, we also dissect branch (lateral) development on the seminal or nodal roots. Genes specific to branch roots have been identified in Arabidopsis (Casimiro et al., 2003; Péret et al., 2009a), rice (Hao and Ichii, 1999; Wang et al., 2006; Zheng et al., 2013), and maize (Hochholdinger and Feix, 1998; Hochholdinger et al., 2001; Woll et al., 2005).This study explored the hypothesis that adult root systems of B. distachyon contain genotypic variation that can be exploited through phenotyping and genotyping to increase cereal yields. A selection of 79 wild lines of B. distachyon from various parts of the Middle East (Fig. 2 shows the geographic origins of the lines) was phenotyped. They were selected for maximum genotypic diversity from 187 diploid lines analyzed with 43 simple sequence repeat markers (Vogel et al., 2009). We phenotyped shoots and mature root systems concurrently because B. distachyon is small enough to complete its life cycle in relatively small pots of soil with minimal influence of pot size compared with crops, such as wheat. We further phenotyped a subset of this population under irrigation (well watered) and drought to assess genotype response to water supply. By conducting whole-plant studies, we aimed to identify phenotypes that described partitioning among shoot and root components and within seed-borne and stem-borne roots. Phenotypes that have the potential to be beneficial to shoot and root components may speed up genetic gain in future.Open in a separate windowFigure 2.B. distachyon lines phenotyped in this study and their geographical origin. Capital letters in parentheses indicate the country of origin: Turkey (T), Spain (S), and Iraq (I; Vogel et al., 2009). a, Adi3, Adi7, Adi10, Adi12, Adi13, and Adi15; b, Bd21 and Bd21-3 are the reference lines of this study. Bd21 was the first sequenced line (Vogel et al., 2010) and root system (described in detail in Watt et al., 2009), and Bd21-3 is the most easily transformed line (Vogel and Hill, 2008) and parent of a T-DNA mutant population (Bragg et al., 2012); c, Gaz1, Gaz4, and Gaz7; d, Kah1, Kah2, and Kah3. e, Koz1, Koz3, and Koz5; f, Tek1 and Tek6; g, exact GPS coordinates are unknown for lines Men2 (S), Mur2 (S), Bd2.3 (I), Bd3-1 (I), and Abr1 (T).  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