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 46 毫秒
1.
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   

2.
Rab Proteins and the Compartmentalization of the Endosomal System     
Angela Wandinger-Ness  Marino Zerial 《Cold Spring Harbor perspectives in biology》2014,6(11)
  相似文献   

3.
The Nuclear Envelope     
Martin W. Hetzer 《Cold Spring Harbor perspectives in biology》2010,2(3)
  相似文献   

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

5.
Noncanonical Sites of Adult Neurogenesis in the Mammalian Brain     
David M. Feliciano  Angélique Bordey  Luca Bonfanti 《Cold Spring Harbor perspectives in biology》2015,7(10)
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   

6.
Cell Biology of Mitotic Recombination     
Michael Lisby  Rodney Rothstein 《Cold Spring Harbor perspectives in biology》2015,7(3)
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   

7.
Archaeology of Eukaryotic DNA Replication     
Kira S. Makarova  Eugene V. Koonin 《Cold Spring Harbor perspectives in biology》2013,5(11)
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   

8.
Mismatch Repair during Homologous and Homeologous Recombination     
Maria Spies  Richard Fishel 《Cold Spring Harbor perspectives in biology》2015,7(3)
  相似文献   

9.
Signaling by Nuclear Receptors     
Richard Sever  Christopher K. Glass 《Cold Spring Harbor perspectives in biology》2013,5(3)
  相似文献   

10.
Orphan Nuclear Bodies     
Maria Carmo-Fonseca  Maria T. Berciano  Miguel Lafarga 《Cold Spring Harbor perspectives in biology》2010,2(9)
  相似文献   

11.
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   

12.
Wnt/Wingless Signaling in Drosophila     
Sharan Swarup  Esther M. Verheyen 《Cold Spring Harbor perspectives in biology》2012,4(6)
  相似文献   

13.
Wnt Pathway Regulation of Embryonic Stem Cell Self-Renewal     
Bradley J. Merrill 《Cold Spring Harbor perspectives in biology》2012,4(9)
Embryonic stem cells (ESCs) can generate all of the cell types found in the adult organism. Remarkably, they retain this ability even after many cell divisions in vitro, as long as the culture conditions prevent differentiation of the cells. Wnt signaling and β-catenin have been shown to cause strong effects on ESCs both in terms of stimulating the expansion of stem cells and stimulating differentiation toward lineage committed cell types. The varied effects of Wnt signaling in ESCs, alongside the sometimes unconventional mechanisms underlying the effects, have generated a fair amount of controversy and intrigue regarding the role of Wnt signaling in pluripotent stem cells. Insights into the mechanisms of Wnt function in stem cells can be gained by examination of the causes for seemingly opposing effects of Wnt signaling on self-renewal versus differentiation.For a single-cell embryo to eventually form an adult organism of trillions of cells, some cells in the early mammalian embryo must be able to generate all cell lineages in the animal. The potential to make all adult cell types defines the property of pluripotency, and it is maintained in proliferating cells through a process called self-renewal. As cells become specified to contribute to particular lineages, they typically lose the ability to make cell types from distinct lineages (Waddington 1957; Hochedlinger and Plath 2009). As such, pluripotency is lost during the initial steps of lineage commitment that occur during gastrulation (Beddington 1982, 1983; Lawson and Pedersen 1987; Lawson et al. 1991), which is a process that coordinates the generation of adult cell lineages with the elaboration of a basic three-dimensional body structure (Heisenberg and Solnica-Krezel 2008). In the mouse, pluripotency can be tested with various experiments; the gold standard is the injection of cells into a blastocyst-staged embryo followed by contribution to a diversity of cell types in the chimeric animal or chimeric embryo after gastrulation. Cells are typically considered to have been pluripotent only if they contributed to all three germ layers (endoderm, mesoderm, and ectoderm).Embryonic stem cells (ESCs) are generated in vitro by outgrowths from a preimplantation-staged embryo, frequently a blastocyst. Pluripotent cells from the inner cell mass (ICM) of the blastocyst proliferate to form colonies, which can be expanded into ESC cultures. When culture conditions for in vitro propagation of mouse ESCs (mESCs) were first discovered more than 30 years ago (Evans and Kaufman 1981; Martin 1981), the critical achievement was finding conditions supporting indefinite ESC self-renewal, that is, maintenance of pluripotency following cell division. Compared with the other cell systems discussed below in this article, mESCs ostensibly display the greatest capacity for self-renewal and the highest ability to maintain pluripotency. As such, mESCs are typically thought to represent a primitive, or “naive,” cellular state in the early embryo.Several culture conditions can support self-renewal of mESCs. Initially, ESCs were grown in serum containing media atop a layer of mitotically inactivated fibroblasts, called feeder cells (Evans and Kaufman 1981). Feeder cells secrete the LIF cytokine, which binds a transmembrane receptor complex consisting of LIFR and gp130 proteins (Gearing et al. 1991; Gearing and Bruce 1992; Davis et al. 1993). LIF binding activates Jak/Stat signaling and Stat3 phosphorylation, which promotes ESC self-renewal (Niwa et al. 1998; Matsuda et al. 1999). Convincing proof of LIF’s importance for self-renewal in vitro was shown when recombinant LIF protein was shown to be sufficient to replace feeder cells in ESC cultures (Smith et al. 1988; Williams et al. 1988; Nichols et al. 1990).Essentially the same feeder cells can be used for both mESCs and human ESCs (hESCs); however, discrete activities of the feeders in terms of the cytokines they release are needed to effect optimal self-renewal for each cell. The LIF cytokine important for mESC self-renewal did not stimulate hESC self-renewal (Thomson et al. 1998). Instead, ERK signaling downstream from Fgf2 must accompany a feeder layer in serum-containing media for optimal hESC self-renewal (Xu et al. 2005). Interestingly, recombinant Fgf2 by itself could not replace feeders, and Fgf2 has been suggested to work in part by stimulating feeders to produce Activin/Nodal ligands; the combination of Fgf2 and Nodal/Activin is sufficient to support hESC self-renewal in serum-free chemically defined culture conditions (Vallier et al. 2004, 2009; James et al. 2005).Clear differences exist between mESCs and hESCs. The colonies adopt different morphologies, they require distinct culture conditions for self-renewal, and they have significantly different gene expression signatures (Brons et al. 2007; Tesar et al. 2007). Mouse EpiSCs are made from the epiblast of postimplantation-staged embryos between embryonic days 5.5 (E5.5) and E6.5 of embryogenesis (Brons et al. 2007; Tesar et al. 2007; Han et al. 2010). Lineage specification of pluripotent epiblast cells begins soon after formation of a cup-like structure, and at E6.5, the cells in the epiblast begin to be specified to primary cell lineages during gastrulation. The in vivo cellular environment for ICM cells and postimplantation epiblast cells is considerably different, and it is not surprising that EpiSCs and mESCs display many different characteristics (Xu et al. 2010). However, it was somewhat surprising that EpiSCs share many characteristics with hESCs, including a common colony morphology, Fgf2 + Activin A culture conditions, and gene expression signatures (Brons et al. 2007; Tesar et al. 2007). Like mESCs and hESCs, EpiSCs pass pluripotency tests for in vitro differentiation and teratoma formation. Whereas mESC can efficiently convert (i.e., differentiate) into EpiSC-like cells when switched to Fgf2/Activin A media (Hanna et al. 2009; Greber et al. 2010), EpiSCs required genetic manipulation or reprogramming for efficient conversion to mES-like cells (Guo et al. 2009; Hanna et al. 2009; Greber et al. 2010; Guo and Smith 2010). Many investigators consider hESCs and mouse EpiSCs to be primed for differentiation as they reside in a less primitive differentiation state relative to the naive state of pluripotency in mESCs.

Table 1.

Pluripotent stem cell states: Naive and primed
Mouse ESCHuman ESCMouse EpiSC
Effects of culture conditions
Serum + Lif
Wnt3a/GSK3inhibitor
Fgf2 + Activin A
Self-renewal
Self-renewal
EpiSC
Differentiation
Differentiation
Self-renewal
Differentiation
Differentiation
Self-renewal
Gene expression profiles
Oct4, Sox2, Nanog
Sox17, Eomes, Fgf5
Klf4, Rex1, Stella
High
Low
High
High
High
Low
High
High
Low
Activity in pluripotency tests
Embryoid body
Teratoma formation
Blastocyst injection
Tetraploid complementation
Pass
Pass
Pass
Pass
Pass
Pass
Not determined
Not determined
Pass
Pass
Poor
Not determined
Epigenetic stateNaivePrimedPrimed
Open in a separate windowThree pluripotent cell systems are compared with respect to characteristics that describe their epigenetic state of pluripotency. See text for details.  相似文献   

14.
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   

15.
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   

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

17.
Kv5, Kv6, Kv8, and Kv9 subunits: No simple silent bystanders     
Elke Bocksteins 《The Journal of general physiology》2016,147(2):105-125
  相似文献   

18.
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   

19.
Focus on Metabolism: Posttranslational Protein Modifications in Plant Metabolism     
Giulia Friso  Klaas J. van Wijk 《Plant physiology》2015,169(3):1469-1487
Posttranslational modifications (PTMs) of proteins greatly expand proteome diversity, increase functionality, and allow for rapid responses, all at relatively low costs for the cell. PTMs play key roles in plants through their impact on signaling, gene expression, protein stability and interactions, and enzyme kinetics. Following a brief discussion of the experimental and bioinformatics challenges of PTM identification, localization, and quantification (occupancy), a concise overview is provided of the major PTMs and their (potential) functional consequences in plants, with emphasis on plant metabolism. Classic examples that illustrate the regulation of plant metabolic enzymes and pathways by PTMs and their cross talk are summarized. Recent large-scale proteomics studies mapped many PTMs to a wide range of metabolic functions. Unraveling of the PTM code, i.e. a predictive understanding of the (combinatorial) consequences of PTMs, is needed to convert this growing wealth of data into an understanding of plant metabolic regulation.The primary amino acid sequence of proteins is defined by the translated mRNA, often followed by N- or C-terminal cleavages for preprocessing, maturation, and/or activation. Proteins can undergo further reversible or irreversible posttranslational modifications (PTMs) of specific amino acid residues. Proteins are directly responsible for the production of plant metabolites because they act as enzymes or as regulators of enzymes. Ultimately, most proteins in a plant cell can affect plant metabolism (e.g. through effects on plant gene expression, cell fate and development, structural support, transport, etc.). Many metabolic enzymes and their regulators undergo a variety of PTMs, possibly resulting in changes in oligomeric state, stabilization/degradation, and (de)activation (Huber and Hardin, 2004), and PTMs can facilitate the optimization of metabolic flux. However, the direct in vivo consequence of a PTM on a metabolic enzyme or pathway is frequently not very clear, in part because it requires measurements of input and output of the reactions, including flux through the enzyme or pathway. This Update will start out with a short overview on the major PTMs observed for each amino acid residue (PTMs, including determination of the localization within proteins (i.e. the specific residues) and occupancy. Challenges in dealing with multiple PTMs per protein and cross talk between PTMs will be briefly outlined. We then describe the major physiological PTMs observed in plants as well as PTMs that are nonenzymatically induced during sample preparation (PTMs, in particular for enzymes in primary metabolism (Calvin cycle, glycolysis, and respiration) and the C4 shuttle accommodating photosynthesis in C4 plants (PTMs observed in plants
Amino Acid ResidueObserved Physiological PTM in PlantsPTMs Caused by Sample Preparation
Ala (A)Not known
Arg (R)Methylation, carbonylation
Asn (N)Deamidation, N-linked gycosylationDeamidation
Asp (D)Phosphorylation (in two-component system)
Cys (C)Glutathionylation (SSG), disulfide bonded (S-S), sulfenylation (-SOH), sulfonylation (-SO3H), acylation, lipidation, acetylation, nitrosylation (SNO), methylation, palmitoylation, phosphorylation (rare)Propionamide
Glu (E)Carboxylation, methylationPyro-Glu
Gln (Q)DeamidationDeamidation, pyro-Glu
Gly (G)N-Myristoylation (N-terminal Gly residue)
His (H)Phosphorylation (infrequent)Oxidation
Ile (I)Not known
Leu (L)Not known
Lys (K)N-ε-Acetylation, methylation, hydroxylation, ubiquitination, sumoylation, deamination, O-glycosylation, carbamylation, carbonylation, formylation
Met (M)(De)formylation, excision (NME), (reversible) oxidation, sulfonation (-SO2), sulfoxation (-SO)Oxidation, 2-oxidation, formylation, carbamylation
Phe (F)Not known
Pro (P)CarbonylationOxidation
Ser (S)Phosphorylation, O-linked glycosylation, O-linked GlcNAc (O-GlcNAc)Formylation
Thr (T)Phosphorylation, O-linked glycosylation, O-linked GlcNAc (O-GlcNAc), carbonylationFormylation
Trp (W)Glycosylation (C-mannosylation)Oxidation
Tyr (Y)Phosphorylation, nitration
Val (V)Not known
Free NH2 of protein N terminiPreprotein processing, Met excision, formylation, pyro-Glu, N-myristoylation, N-acylation (i.e. palmitoylation), N-terminal α-amine acetylation, ubiquitinationFormylation (Met), pyro-Glu (Gln)
Open in a separate window

Table II.

Most significant and/or frequent PTMs observed in plants
Type of PTM (Reversible, Except if Marked with an Asterisk)Spontaneous (S; Nonenzymatic) or Enzymatic (E)Comment on Subcellular Location and Frequency
Phosphorylation (Ser, Thr, Tyr, His, Asp)EHis and Asp phosphorylation have low frequency
S-Nitrosylation (Cys) and nitration* (Tyr)S (RNS), but reversal is enzymatic for Cys by thioredoxinsThroughout the cell
Acetylation (N-terminal α-amine, Lys ε-amine)EIn mitochondria, very little N-terminal acetylation, but high Lys acetylation; Lys acetylation correlates to [acetyl-CoA]
Deamidation (Gln, Asn)S, but reversal of isoAsp is enzymatic by isoAsp methyltransferaseThroughout the cell
Lipidation (S-acetylation, N-meristoylation*, prenylation*; Cys, Gly, Lys, Trp, N terminal)ENot (or rarely) within plastids, mitochondria, peroxisomes
N-Linked glycosylation (Asp); O linked (Lys, Ser, Thr, Trp)EOnly proteins passing through the secretory system; O linked in the cell wall
Ubiquination (Lys, N terminal)ENot within plastids, mitochondria, peroxisomes
Sumoylation (Lys)ENot within plastids, mitochondria, peroxisomes
Carbonylation* (Pro, Lys, Arg, Thr)S (ROS)High levels in mitochondria and chloroplast
Methylation (Arg, Lys, N terminal)EHistones (nucleus) and chloroplasts; still underexplored
Glutathionylation (Cys)EHigh levels in chloroplasts
Oxidation (Met, Cys)S (ROS) and E (by PCOs; see Fig. 1B), but reversal is enzymatic by Met sulfoxide reductases, glutaredoxins, and thioredoxins, except if double oxidizedHigh levels in mitochondria and chloroplast
Peptidase* (cleavage peptidyl bond)EThroughout the cell
S-Guanylation (Cys)S (RNS)Rare; 8-nitro-cGMP is signaling molecule in guard cells
Formylation (Met)S, but deformylation is enzymatic by peptide deformylaseAll chloroplasts and mitochondria-encoded proteins are synthesized with initiating formylated Met
Open in a separate window

Table III.

Regulation by PTMs in plant metabolism and classic examples of well-studied enzymes and pathwaysMany of these enzymes also undergo allosteric regulation through cellular metabolites. GAPDH, Glyceraldehyde-3-phosphate dehydrogenase; PRK, phosphoribulokinase.
ProcessEnzymesPTMs, Protein Modifiers, LocalizationReferences
Calvin-Benson cycle (chloroplasts)Many enzymesOxidoreduction of S-S bonds, reversible nitrosylation, glutathionylation; through ferredoxin/ferredoxin-thioredoxin reductase/thioredoxins (mostly f and m) and glutaredoxins; proteomics studies in Arabidopsis and C. reinhardtiiMichelet et al. (2013)
RubiscoMethylation, carbamylation, acetylation, N-terminal processing, oligomerization; classical studies in pea (Pisum sativum), spinach (Spinacia oleracea), and ArabidopsisHoutz and Portis (2003); Houtz et al. (2008)
GAPDH/CP12/PRK supercomplexDynamic heterooligomerization through reversible S-S bond formation controlled by thioredoxinsGraciet et al. (2004); Michelet et al. (2013); López-Calcagno et al. (2014)
GlycolysisCytosolic PEPCPhosphorylation (S, T), monoubiquitinationO’Leary et al. (2011)
PhotorespirationSeven enzymes are phosphorylatedPhosphorylation from meta-analysis of public phosphoproteomics data for Arabidopsis; located in chloroplasts, peroxisomes, mitochondriaHodges et al. (2013)
Maize glycerate kinaseRedox-regulated S-S bond; thioredoxin f; studied extensively in chloroplasts of C4 maizeBartsch et al. (2010)
Respiration (mitochondria)Potentially many enzymes, but functional/biochemical consequences are relatively unexploredRecent studies suggested PTMs for many tricarboxylic acid cycle enzymes, including Lys acetylation and thioredoxin-driven S-S formation; in particular, succinate dehydrogenase and fumarase are inactivated by thioredoxinsLázaro et al. (2013); Schmidtmann et al. (2014); Daloso et al. (2015)
PDHSer (de)phosphorylation by intrinsic kinase and phosphatase; ammonia and pyruvate control PDH kinase activity; see Figure 1BThelen et al. (2000); Tovar-Méndez et al. (2003)
C4 cycle (C3 and C4 homologs also involved in glycolysis and/or gluconeogenesis)Pyruvate orthophosphate dikinasePhosphorylation by pyruvate orthophosphate dikinase-RP, an S/T bifunctional kinase-phosphatase; in chloroplastsChastain et al. (2011); Chen et al. (2014)
PEPCPhosphorylation; allosteric regulation by malate and Glc-6-P; in cytosol in mesophyll cells in C4 species (e.g. Panicum maximum); see Figure 1AIzui et al. (2004); Bailey et al. (2007)
PEPC kinaseUbiquitination resulting in degradation (note also diurnal mRNA levels and linkage to activity level; very low protein level); in cytosol in mesophyll cells in C4 species (e.g. Flaveria spp. and maize)Agetsuma et al. (2005)
PEPC kinasePhosphorylation in cytosol in bundle sheath cellsBailey et al. (2007)
Starch metabolism (chloroplasts)ADP-Glc pyrophosphorylaseRedox-regulated disulfide bonds and dynamic oligomerization; thioredoxins; see Figure 1CGeigenberger et al. (2005); Geigenberger (2011)
Starch-branching enzyme IIPhosphorylation by Ca2+-dependent protein kinase; P-driven heterooligomerizationGrimaud et al. (2008); Tetlow and Emes (2014)
Suc metabolism (cytosol)SPS (synthesis of Suc)(De)phosphorylation; SPS kinase and SPS phosphatase; 14-3-3 proteins; cytosol (maize and others)Huber (2007)
Suc synthase (breakdown of Suc)Phosphorylation; Ca2+-dependent protein kinase; correlations to activity, localization, and turnoverDuncan and Huber (2007); Fedosejevs et al. (2014)
Photosynthetic electron transport (chloroplast thylakoid membranes)PSII core and light-harvesting complex proteins(De)phosphorylation by state-transition kinases (STN7/8) and PP2C phosphatases (PBCP and PPH1/TAP38)Pesaresi et al. (2011); Tikkanen et al. (2012); Rochaix (2014)
Nitrogen assimilationNitrate reductase(De)phosphorylation; 14-3-3 proteinsLillo et al. (2004); Huber (2007)
Open in a separate windowThere are many recent reviews focusing on specific PTMs in plant biology, many of which are cited in this Update. However, the last general review on plant PTMs is from 2010 (Ytterberg and Jensen, 2010); given the enormous progress in PTM research in plants over the last 5 years, a comprehensive overview is overdue. Finally, this Update does not review allosteric regulation by metabolites or other types of metabolic feedback and flux control, even if this is extremely important in the regulation of metabolism and (de)activation of enzymes. Recent reviews for specific pathways, such as isoprenoid metabolism (Kötting et al., 2010; Banerjee and Sharkey, 2014; Rodríguez-Concepción and Boronat, 2015), tetrapyrrole metabolism (Brzezowski et al., 2015), the Calvin-Benson cycle (Michelet et al., 2013), starch metabolism (Kötting et al., 2010; Geigenberger, 2011; Tetlow and Emes, 2014), and photorespiration (Hodges et al., 2013) provide more in-depth discussions of metabolic regulation through various posttranslational mechanisms. Many of the PTMs that have been discovered in the last decade through large-scale proteomics approaches have not yet been integrated in such pathway-specific reviews, because these data are not always easily accessible and because the biological significance of many PTMs is simply not yet understood. We hope that this Update will increase the general awareness of the existence of these PTM data sets, such that their biological significance can be tested and incorporated in metabolic pathways.  相似文献   

20.
Conflict on the Sex Chromosomes: Cause,Effect, and Complexity     
Judith E. Mank  David J. Hosken  Nina Wedell 《Cold Spring Harbor perspectives in biology》2014,6(12)
Intralocus sexual conflict and intragenomic conflict both affect sex chromosome evolution and can in extreme cases even cause the complete turnover of sex chromosomes. Additionally, established sex chromosomes often become the focus of heightened conflict. This creates a tangled relationship between sex chromosomes and conflict with respect to cause and effect. To further complicate matters, sexual and intragenomic conflict may exacerbate one another and thereby further fuel sex chromosome change. Different magnitudes and foci of conflict offer potential explanations for lineage-specific variation in sex chromosome evolution and answer long-standing questions as to why some sex chromosomes are remarkably stable, whereas others show rapid rates of evolutionary change.Compared to the autosomes, the unique inheritance pattern of the sex chromosomes is often thought to intensify evolutionary conflict (Rice 1984; Frank 1991; Jaenike 2001). Sex chromosomes are therefore hot spots for two specific types of conflict: intralocus conflict, in which an allele confers different fitness effects depending on the sex in which it is found, and intragenomic conflict, in which selfish genetic elements (SGEs) promote their own transmission at the expense of unlinked regions of the genome. These conflicts act in distinct but complementary ways. Not only do they shape sex chromosome and genome evolution, but in some cases, they also have the power to cause complete turnover of sex chromosomes.Unlike the autosomes, because the sex chromosomes are unevenly transmitted between males and females and are also unevenly distributed between the sexes (Fig. 1), the relative effect of male- and female-specific selection acting on them is unbalanced. The inherent differences in sex-specific selection on the sex chromosomes themselves, and between the sex chromosomes and the autosomes, form the basis of a large and often compelling body of evolutionary theory that predicts the ways that intralocus sexual conflict will arise, play out, and in some cases potentially be resolved. This theory predicts that, under some conditions, the sex chromosomes are hot spots of intralocus sexual conflict (Rice 1984; Albert and Otto 2005; Connallon and Clark 2010), and in some cases alleles that harm one sex more than they benefit the other can still reach high frequencies if they are sex-linked (Rice 1984; Dean et al. 2012). All this theory predicts that although the sex chromosomes generally represent a small proportion of the genome, they should play a disproportionately large role in sexual conflict, sexual dimorphism, and sexual selection. There is substantial empirical evidence supporting at least some of this theory (Dean and Mank 2014).Open in a separate windowFigure 1.Transmission of the sex chromosomes. Females are shown in red, males in blue. In male heterogamety (A), the Y chromosome is passed through the patriline and limited to males. The maternal X chromosome is passed from mother to both sons and daughters, but the paternal X can only be transmitted to daughters. Additionally, the X is present two-thirds of the time in females. In female heterogamety (B), the W chromosome is limited to females and passed solely from mother to daughter. The paternal Z chromosome can be passed from father to both daughters and sons; however, the maternal Z chromosome is only passed to sons. Converse to the X chromosome, the Z chromosome is resident in males two-thirds of the time.

Table 1.

Studies showing a disproportionate role of the sex chromosomes in sexual dimorphism, fitness, or fertility
Male heterogametyFemale heterogamety
X chromosomeY chromosomeZ chromosomeW chromosome
Associations with sexually dimorphic phenotypic traitsSize dimorphism in red deer (Foerster et al. 2007)
Sexually antagonistic fitness variation in Drosophila (Gibson et al. 2002; Innocenti and Morrow 2010)
Variation in stalk-eyed fly eye span (Wolfenbarger and Wilkinson 2001)
Male coloration in guppies (Winge 1927; Postma et al. 2011)
Male mating behavior in sticklebacks (Kitano et al. 2009; although see Natri et al. 2013)
Female mating preference in moths (Iyengar et al. 2002)
Female mate choice in flycatchers (Saether et al. 2007)
Male plumage traits in flycatchers (Saetre et al. 2003)
Female benefit coloration in cichlids (Roberts et al. 2009)
Associations with reproduction and fertilityMale reproductive genes in mice (Mueller et al. 2008)Male fertility genes in mammals (Lange et al. 2009)
Male fitness in Drosophila (Lemos et al. 2010; Sackton et al. 2011)
Genes related to male reproductive function in silk moth (Arunkumar et al. 2009)Genes related to female fecundity and fertility in chickens (Moghadam et al. 2012)
Open in a separate windowIn addition to their role in sexual conflict, the sex chromosomes also experience high levels of intragenomic conflict resulting from SGEs that promote their own transmission at the expense of the rest of the genome (Burt and Trivers 2006). Of particular importance are sex-linked segregation distorters, as these generate strong selection favoring genes that suppress their action. Theory predicts that intragenomic conflict should be particularly intense when involving the sex chromosomes (Hurst and Pomiankowski 1991). This is because sex chromosomes can generate antagonistic coevolution between sex-linked segregation distorters and their suppressors (Partridge and Hurst 1998).In addition to shaping the evolutionary properties of existing sex chromosomes, in some situations sexual and intragenomic conflict may also actually catalyze the formation of sex chromosomes (Fisher 1931; Charlesworth and Charlesworth 1980; Rice 1987; Werren and Beukeboom 1998). Conflict may also explain the rate of degradation of the sex-limited Y and W chromosomes (Bachtrog et al. 2011) and turnover of sex chromosomes (Hall 2004; van Doorn and Kirkpatrick 2007, 2010), suggesting that conflict plays a causal role in sex chromosome evolution.There is also growing evidence of a direct relationship between SGEs and sexual conflict involving the sex chromosomes. Many SGEs increase their transmission advantage by targeting sperm, which can reduce male fertility owing to reduced overall sperm production. Lower sperm production can in turn result in reduced siring success during sperm competition (Price and Wedell 2008) and potentially favor polyandry as a female strategy to bias paternity against SGE-carrying males (Wedell 2013). As a consequence, SGEs may also influence the potential for sexual conflict as they can favor increased female mating rates, which increases the potential for conflict between the sexes.Sexual and intragenomic conflict are therefore critical for sex chromosome evolution and, once sex chromosomes are established, can further shape their evolutionary and genomic properties. Disentangling cause and effect is difficult but crucial to understanding the role of sex chromosomes in sexual and intragenomic conflict and vice versa.  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