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 859 毫秒
1.
2.
A developing animal is exposed to both intrinsic and extrinsic stresses. One stress response is caspase activation. Caspase activation not only controls apoptosis but also proliferation, differentiation, cell shape, and cell migration. Caspase activation drives development by executing cell death or nonapoptotic functions in a cell-autonomous manner, and by secreting signaling molecules or generating mechanical forces, in a noncell autonomous manner.Programmed cell death or apoptosis occurs widely during development. During C. elegans development, 131 cells die by caspase CED-3-dependent apoptosis; however, ced-3 mutants do not show significant developmental defects (Ellis and Horvitz 1986). In contrast, studies on caspase mutants in mouse and Drosophila have revealed caspases’ roles in development. During development, cells are exposed to extrinsic and intrinsic stresses, and caspases are activated as one of multiple stress responses that ensure developmental robustness (Fig. 1). Caspases actively regulate animal development through both apoptosis and nonapoptotic functions that involve cell–cell communication in developing cell communities (Miura 2011). This chapter focuses on the in vivo roles of caspases in development and regeneration.Open in a separate windowFigure 1.Caspase activation during development. An embryo undergoes intrinsic and extrinsic stress, which activates caspases to execute both apoptotic and nonapoptotic functions, including cell differentiation and dendrite pruning. Apoptotic cells affect the shape and behavior of their neighboring cells. Caspase-activated cells are shown in dark gray.  相似文献   

3.
Gong B  Wang YT 《The EMBO journal》2012,31(4):783-785
EMBO J 31 4, 805–816 (2012); published online December202011Synaptic plasticity, the activity-dependent modification of synaptic strength, plays a fundamental role in learning and memory as well as in developmental maturation of neuronal circuitry. However, how synaptic plasticity is induced and regulated remains poorly understood. In this issue of The EMBO Journal, Yang and colleagues present sets of exciting data, suggesting that G-protein-coupled receptors (GPCRs) selectively execute distinct signalling pathways to differentially regulate induction thresholds of hippocampal long-term potentiation (LTP) and long-term depression (LTD), thereby governing the direction of synaptic plasticity. These results shed significant light on our current understanding of how bidirectional synaptic plasticity is regulated.Synaptic plasticity has been demonstrated at synapses in various brain regions; the most well-characterized forms are LTP and LTD at hippocampal CA1 glutamatergic synapses (Collingridge et al, 2004). In experimental models, LTP and LTD can be, respectively, induced by high-frequency stimulation (HFS) and low-frequency stimulation (LFS) via activation of the N-methyl-D-aspartic acid (NMDA) subtype ionotropic glutamate receptor (NMDAR). However, how HFS and LFS activate NMDARs and thereby lead to synaptic plasticity remains poorly understood and highly controversial. It is even more unclear how the bidirectional synaptic plasticity is produced and regulated in response to physiological or pathological changes.Functional NMDARs consist primarily of two GluN1 subunits and two GluN2 subunits, with GluN2A and GluN2B subunits being the most common NMDAR subunits found in the cortical and hippocampal regions of the adult brain (Cull-Candy et al, 2001). GluN2A and GluN2B subunits may confer distinct gating and pharmacological properties to NMDARs and couple them to distinct intracellular signalling machineries (Cull-Candy et al, 2001). Moreover, the ratio of these two subpopulations of NMDARs at the glutamatergic synapse is dynamically regulated in an activity-dependent manner (Bellone and Nicoll, 2007; Cho et al, 2009; Xu et al, 2009). Although controversial, GluN2A- and GluN2B-containing NMDARs have been suggested to have differential roles in regulating the direction of synaptic plasticity (Collingridge et al, 2004; Morishita et al, 2007). Among the factors shown to regulate NMDAR function, Src family tyrosine kinases may be the best characterized, with both Src and Fyn able to upregulate NMDAR function, and thus LTP induction (Salter and Kalia, 2004). However, if these kinases modulate NMDAR function in a NMDAR subunit-specific manner remains unknown. To explore this concept, Yang et al (2012) investigated the potential subunit-specific regulation of NMDARs by Src and Fyn using whole-cell patch clamp recording of NMDAR-mediated currents from acutely dissociated CA1 hippocampal neurons or from rat hippocampal slices. They found that intracellular perfusion of recombinant Src or Fyn increased the NMDAR-mediated currents. By applying subunit-preferential antagonists of GluN2A- or GluN2B-containing NMDARs, or by using neurons obtained from GluN2A knockout mice, they discovered that Src and Fyn differentially enhanced currents gated through GluN2A- and GluN2B-containing NMDARs, respectively.Can physiological or pathological factors differentially activate Src or Fyn, thereby exerting subunit-specific regulation of NMDAR function? To answer this question, Yang et al focused their investigation on the role of GPCRs, specifically pituitary adenylate cyclase activating peptide receptor (PAC1R) and dopamine D1 receptor (D1R), both of which have recently been shown to potentiate NMDARs through Src family kinases (Macdonald et al, 2005; Hu et al, 2010). Indeed, they found that activation of PAC1R specifically increased GluN2A-NMDAR-mediated currents without affecting currents gated through GluN2B-NMDARs, and this potentiation was prevented by the Src-specific inhibitory peptide Src(40–58) (Salter and Kalia, 2004). To rule out the contribution of Fyn, the authors developed a novel-specific Fyn inhibitory peptide Fyn(39–57), and demonstrated that it had little effect on PAC1R potentiation. In contrast, activation of D1R potentiated GluN2B- (but not GluN2A-) NMDAR-mediated currents, and this potentiation was specifically eliminated by Fyn(39–57), but not by Src(40–58). The authors further demonstrated that stimulation of PAC1Rs resulted in a selective activation of Src kinase and consequent tyrosine phosphorylation of the GluN2A subunit, whereas activation of D1Rs led to a specific increase in Fyn-mediated tyrosine phosphorylation of the GluN2B subunit. To provide convincing evidence that these subunit-differential modulations are indeed the result of tyrosine phosphorylation of the respective NMDAR subunits, the authors then performed electrophysiological experiments using neurons from two knockin mouse lines GluN2A(Y1325F) and GluN2B(Y1472F), in which the tyrosine phosphorylation residues in native GluN2A and GluN2B subunits were, respectively, replaced with non-phosphorylatable phenylalanine residues. As expected, the authors found that PAC1R-mediated potentiation of NMDA currents was lost in neurons from GluN2A(Y1325F) mice (but maintained in neurons from GluN2B(Y1472F) mice), while D1R-mediated enhancement of NMDA currents was only observed in neurons from GluN2A(Y1325F) mice. Together, as illustrated in Figure 1, the authors have made a very convincing case that PAC1R and D1R, respectively, enhance function of GluN2A- and GluN2B-containing NMDARs by differentially activating Src- and Fyn-mediated phosphorylation of respective NMDAR subunits.Open in a separate windowFigure 1GPCRs regulate the direction of synaptic plasticity via activating distinct signalling pathways. Synaptic NMDA receptors, both GluN2A- and GluN2B-containing, play key roles in the induction of various forms of synaptic plasticity at the hippocampal CA1 glutamatergic synapse. Under the basal level of GluN2A and GluN2B ratio, stimulation with a train of pulses at frequencies from 1 to 100 Hz produces a frequency and plasticity (LTD–LTP) curve, with maximum LTD and LTP being, respectively, induced at 1 and 100 Hz. Activation of PAC1R with its agonist PACAP38 activates Src and thereby results in tyrosine phosphorylation and consequent functional upregulation of GluN2A-containing NMDARs, resulting in an increase in the ratio of functional GluN2A and GluN2B. The increased ratio in turn causes a left shift of frequency and plasticity curve, favouring LTP induction. In contrast, activation of D1R by the receptor agonist SKF81297 triggers Fyn-specific tyrosine phosphorylation and functional upregulation of GluN2B, causing a reduction of GluN2A and GluN2B ratio. This decreased ratio results in a right shift of the curve, favouring LTD induction. The ability of GPCRs to differentially activate distinct downstream signalling pathways involved in synaptic plasticity suggests the potential roles of GPCRs in governing the direction of synaptic plasticity.Given the coupling of NMDARs to the induction of synaptic plasticity, it is then reasonable to ask if activation of the two GPCRs can selectively affect the induction of LTP or LTD at CA1 synapses. Yang and colleagues investigated the effects of pharmacological activation of PAC1R and D1R on the induction of LTP and LTD by recording the field excitatory postsynaptic potentials from hippocampal slices. Consistent with differential roles of NMDAR subunits in governing directions of synaptic plasticity, the authors observed that activation of PAC1Rs reduces the induction threshold of LTP, while stimulation of D1Rs favours LTD induction (Figure 1). Facilitation of LTP by PAC1R and LTD by D1R were, respectively, prevented in the brain slices obtained from GluN2A(Y1325F) and GluN2B(Y1472F) knockin mice, supporting the differential involvements of Src-mediated GluN2A phosphorylation and Fyn-mediated GluN2B phosphorylation.Taken together, the authors'' results have demonstrated that activation of PAC1R and D1R can control the direction of synaptic plasticity at the hippocampal CA1 synapse by differentially regulating NMDAREPSCs in a subunit-specific fashion (Figure 1). Specifically, PAC1R enhances the function of GluN2A-containing NMDARs by increasing Src phosphorylation of GluN2A subunit at Y1325, whereas D1R upregulates GluN2B-containing NMDARs through increased Fyn phosphorylation of GluN2B at Y1472. Moreover, by regulating the ratio of functional GluN2A- and GluN2B-containing NMDARs, PAC1R and D1R in turn modulate the direction of synaptic plasticity, favouring the production of LTP and LTD, respectively.While consistent with the recently proposed hypothesis that GluN2A and GluN2B may have preferential roles in the induction of hippocampal CA1 LTP and LTD (Collingridge et al, 2004; but see also Morishita et al, 2007), the current study further emphasizes the importance of GluN2A/GluN2B ratios in regulating LTP and LTD thresholds: increased ratio favours LTP, while reduced ratio promotes LTD. However, this seems to contradict some recent studies where the reduction and increase in the GluN2A/GluN2B ratio appeared to, respectively, favour LTP (Cho et al, 2009; Xu et al, 2009) and LTD (Xu et al, 2009). Therefore, the direction of plasticity change is likely modulated not only by the GluN2A/GluN2B ratio, but also by additional factors such as experimental conditions, developmental stages, and brain regions.Under many experimental conditions, LTP and LTD are usually induced by HFS and LFS stimulating protocols, respectively, but it remains essentially unknown how LTP and LTD are physiologically or pathologically generated in animals. To this end, the identification of different GPCRs as the endogenous upstream regulators of NMDA receptor subpopulations, and hence regulators of synaptic plasticity, is the major novelty of Yang and colleagues'' work. Future studies are needed to investigate if and how PAC1R and/or D1R are critically involved in the production of LTP or LTD in animals under physiological or pathological conditions. Given the fact that Src family kinases may be required for LTP induced by HFS in hippocampal slices (Salter and Kalia, 2004), an equally intriguing question would be whether these GPCRs are actually required for LTP/LTD induced by HFS/LFS experimental paradigms. In line with this conjecture, it would be interesting to determine if ligands for various GPCRs co-exist in the glutamatergic presynaptic terminals and, if so, can be differentially co-released with glutamate in a frequency-dependent manner, thereby contributing to either HFS-induced LTP or LFS-induced LTD.The findings by Yang and colleagues establish an exciting mechanistic model by which GPCRs can govern the direction of synaptic plasticity by determining the contributions of GluN2A- and GluN2B-NMDARs through differential tyrosine phosphorylation of respective NMDA receptor subtypes. Additional studies further validating this model under physiological and pathological conditions will greatly improve our understanding of the molecular mechanisms underlying synaptic plasticity and cognitive brain functions. In addition, NMDARs, depending on their subunit composition and/or subcellular localization, may also have complex roles in mediating neuronal survival and death (Lai et al, 2011). Considering that neurotoxicity produced by over-activation of NMDARs is widely accepted to be a common mechanism for neuronal loss in a number of acute brain injuries and chronic neurodegenerative diseases, Yang and colleagues'' finding of the differential regulation of NMDAR subunits by different GPCRs could have wider implications beyond synaptic plasticity.  相似文献   

4.
5.
The primary goal of mitosis is to partition duplicated chromosomes into daughter cells. Eukaryotic chromosomes are equipped with two distinct classes of intrinsic machineries, cohesin and condensins, that ensure their faithful segregation during mitosis. Cohesin holds sister chromatids together immediately after their synthesis during S phase until the establishment of bipolar attachments to the mitotic spindle in metaphase. Condensins, on the other hand, attempt to “resolve” sister chromatids by counteracting cohesin. The products of the balancing acts of cohesin and condensins are metaphase chromosomes, in which two rod-shaped chromatids are connected primarily at the centromere. In anaphase, this connection is released by the action of separase that proteolytically cleaves the remaining population of cohesin. Recent studies uncover how this series of events might be mechanistically coupled with each other and intricately regulated by a number of regulatory factors.In eukaryotic cells, genomic DNA is packaged into chromatin and stored in the cell nucleus, in which essential chromosomal processes, including DNA replication and gene expression, take place (Fig. 1, interphase). At the onset of mitosis, the nuclear envelope breaks down and chromatin is progressively converted into a discrete set of rod-shaped structures known as metaphase chromosomes (Fig. 1, metaphase). In each chromosome, a pair of sister kinetochores assembles at its centromeric region, and their bioriented attachment to the mitotic spindle acts as a prerequisite for equal segregation of sister chromatids. The linkage between sister chromatids is dissolved at the onset of anaphase, allowing them to be pulled apart to opposite poles of the cell (Fig. 1, anaphase). At the end of mitosis, the nuclear envelope reassembles around two sets of segregated chromatids, leading to the production of genetically identical daughter cells (Fig. 1, telophase).Open in a separate windowFigure 1.Overview of chromosome dynamics during mitosis. In addition to the crucial role of kinetochore–spindle interactions, an intricate balance between cohesive and resolving forces acting on sister chromatid arms (top left, inset) underlies the process of chromosome segregation. See the text for major events in chromosome segregation.Although the centromere–kinetochore region plays a crucial role in the segregation process, sister chromatid arms also undergo dynamic structural changes to facilitate their own separation. Conceptually, such structural changes are an outcome of two balancing forces, namely, cohesive and resolving forces (Fig. 1, top left, inset). The cohesive force holds a pair of duplicated arms until proper timing of separation, otherwise daughter cells would receive too many or too few copies of chromosomes. The resolving force, on the other hand, counteracts the cohesive force, reorganizing each chromosome into a pair of rod-shaped chromatids. From this standpoint, the pathway of chromosome segregation is regarded as a dynamic process, in which the initially robust cohesive force is gradually weakened and eventually dominated by the resolving force. Almost two decades ago, genetic and biochemical studies for the behavior of mitotic chromosomes converged productively, culminating in the discovery of cohesin (Guacci et al. 1997; Michaelis et al. 1997; Losada et al. 1998) and condensin (Hirano et al. 1997; Sutani et al. 1999), which are responsible for the cohesive and resolving forces, respectively. The subsequent characterizations of these two protein complexes have not only transformed our molecular understanding of chromosome dynamics during mitosis and meiosis, but also provided far-reaching implications in genome stability, as well as unexpected links to human diseases. In this article, I summarize recent progress in our understanding of mitotic chromosome dynamics with a major focus on the regulatory networks surrounding cohesin and condensin. I also discuss emerging topics and attempt to clarify outstanding questions in the field.  相似文献   

6.
EMBO J (2013) 32: 496–510 doi:10.1038/emboj.2012.334; published online January042013Alteration of the efficacy of excitatory synaptic transmission between neurons is a critical element in the processes of learning, memory, and behaviour. Despite decades of research aimed at elucidating basic cellular mechanisms underlying synaptic plasticity, new pathways and permutations continue to be discovered. Carta et al (2013) now show that activation of the calcium/calmodulin dependent kinase II (CaMKII) induces an unusual postsynaptic form of long-term depression (LTD) at the hippocampal mossy fibre synapse by promoting lateral diffusion of kainate receptors (KARs), a family of ionotropic glutamate receptors (iGluRs) that influence pyramidal neuron excitability. This report therefore reveals a new and mechanistically unique way of fine-tuning synaptic plasticity at this central synapse in the hippocampus.Information transfer within the nervous system is regulated at the synaptic level by diverse cellular mechanisms. Synaptic efficacy is not static (i.e., it is ‘plastic''), and the capacity to adjust the strength of communication between neurons in a network has been shown to be a critical component of diverse aspects of brain function that include many forms of behavioural learning (Martin et al, 2000). The complex means by which neurons adjust their synaptic properties in response to changes in local and global activity in the central nervous system has been the subject of intensive investigation spanning multiple decades (Malenka and Bear, 2004; Feldman, 2009). Nonetheless, new mechanisms underlying plasticity of excitatory and inhibitory synaptic transmission continue to be elucidated; these can vary depending on the experimental parameters for induction of plasticity, the particular type of synapse under investigation, and even the prior history of activation at the synapse. Long-term potentiation (LTP) and LTD of excitatory synaptic transmission are two well-known phenomena in which efficacy is increased or decreased, respectively, and at many synapses in the CNS occur through concomitant alterations in the number of postsynaptic iGluRs. The movement of excitatory receptors in and out of synapses, and more generally to and from the neuronal plasma membrane, is dictated by their association with a wide variety of scaffolding and chaperone proteins, whose interactions are often controlled by various protein kinases (Anggono and Huganir, 2012).It is generally appreciated now that long-term synaptic plasticity can be elicited by a variety of mechanisms even within a single type of synaptic connection. In addition to postsynaptic alterations in receptor content, for example, synaptic efficacy can also be tuned by regulated alterations in the probability of vesicular release of the neurotransmitter. Until recently, this presynaptic form of plasticity was thought to be the exclusive mechanism for altering excitatory synaptic strength at a morphologically unusual synapse in the hippocampus formed between large bouton-like presynaptic terminals arising from granule cell axons, or mossy fibres, and proximal dendrites on CA3 pyramidal neurons (Nicoll and Schmitz, 2005). These synaptic connections allow for single dentate granule cells to profoundly influence the likelihood of action potential firing in CA3 pyramidal neurons in a frequency-dependent manner, and for that reason have been referred to as ‘conditional detonator'' synapses (Henze et al, 2002). The precise mechanisms that lead to increased vesicular release probability following LTP-inducing stimulation of mossy fibre axons, including a potential role for retrograde signalling, remain the subject of debate, although there is general consensus that activation of presynaptic protein kinase A (PKA) is a key step in this form of synaptic plasticity (Figure 1A). Enhancing release probability impacts signalling through all three types of iGluRs present at mossy fibre synapses—AMPA, NMDA, and KARs. Recently, however, novel postsynaptic forms of mossy fibre plasticity were discovered in which induction protocols specifically increased the number of NMDA receptors (Kwon and Castillo, 2008; Rebola et al, 2008) or decreased the number of KARs (Selak et al, 2009), expanding the mechanistic repertoire at this historical site of focus of research on presynaptic LTP. Alterations in the synaptic content of particular iGluRs could serve as an additional means to fine-tune synaptic integration at the mossy fibre—CA3 synapse and therefore have important consequences for hippocampal network excitability.Open in a separate windowFigure 1Kainate receptor-dependent plasticity mechanisms at the hippocampal mossy fibre–CA3 synapse. (A) Activation of presynaptic receptors enhances glutamate release from the mossy fibre terminals. (B) A spike-timing-dependent plasticity protocol known to activate postsynaptic CaMKII results in long-term synaptic depression. CaMKII phosphorylates the GluK5 kainate receptor subunit, which uncouples the receptor from PSD-95 in the postsynaptic density. This leads to an increase in receptor mobility and diffusion away from the synapse. (C) Low-frequency stimulation of mossy fibres and activation of postsynaptic group 1 mGluRs leads to activation of PKC, which promotes the association of SNAP-25 to the GluK5 kainate receptor subunit and the subsequent endocytosis of synaptic receptors.In this issue, Carta et al (2013) identify a new postsynaptic mechanism for shaping mossy fibre plasticity that is specific to synaptic KARs, which serve to influence temporal integration of synaptic input as well as pyramidal neuron excitability through modulation of intrinsic ion channels. The authors paired postsynaptic depolarization of CA3 pyramidal neurons with a precisely timed presynaptic release of glutamate in a pattern that is known to produce LTP at many central synapses (Feldman, 2012). At mossy fibre synapses, however, this form of spike-timing-dependent plasticity (STDP) instead caused LTD of KAR-mediated excitatory synaptic potentials (KAR-LTD) while leaving AMPA receptor function unaltered (Figure 1B) (Carta et al, 2013). Using a series of genetic and pharmacological manipulations, Carta et al (2013) found that KAR-LTD was dependent upon the activation of postsynaptic KARs themselves, a rise in postsynaptic Ca2+, and CaMKII phosphorylation of a specific protein component of synaptic KARs, the GluK5 subunit. Unlike other mechanisms of postsynaptic mossy fibre plasticity, KAR-LTD was independent of NMDA or metabotropic glutamate receptor activation. Most surprisingly, KAR-LTD did not require receptor endocytosis from the plasma membrane, as is the case with most other forms of postsynaptic depression of excitatory transmission, including a distinct form of KAR-LTD reported previously (Selak et al, 2009) (Figure 1C). Instead, CaMKII-mediated phosphorylation of GluK5 subunits likely uncoupled receptors from the postsynaptic scaffolding protein PSD-95, which then led to enhanced lateral diffusion of KARs out of mossy fibre synapses. As KAR endocytosis was not altered in mossy fibre STDP, the activity-dependent reduction in KAR signalling was effectively limited to those receptors in the synapse. A molecular replacement strategy was employed using biolistic-based expression of mutant KARs in cultured hippocampal slices prepared from KAR knockout mice, which allowed Carta et al (2013) to corroborate their detailed biochemical studies by showing that reconstituted KAR currents in CA3 neurons expressing recombinant GluK5 phosphorylation site substitutions were unable to express KAR-LTD. In summary, KAR-mediated activation of CaMKII leads to phosphorylation of the GluK5 subunit and subsequent KAR-LTD through enhanced lateral mobility of synaptic receptors (Figure 1B).These findings are intriguing for several reasons. Most notably, they stand in stark contrast to studies in which CaMKII activation primarily triggers potentiation, rather than depression, of excitatory synaptic transmission at other synapses (Lisman et al, 2012). CaMKII recently was shown to cause diffusional trapping of AMPA receptor complexes within the postsynaptic density following phosphorylation of a closely associated auxiliary subunit, stargazin (Opazo et al, 2010), which is precisely the opposite of the effects of activation of the enzyme on KAR mobility at mossy fibre synapses. Further, these divergent consequences are both dependent upon carboxy-terminal PDZ interactions with scaffolding proteins, although in each case further research is needed to dissect out the relevant binding partners that control lateral mobility. It is of interest that KAR-LTD required synaptic activation of KARs to initiate signalling via CaMKII, which implies a tight coupling exists between KARs and the holoenzyme in the mossy fibre postsynaptic density. This observation also raises the possibility that activated CaMKII could phosphorylate other targets to effect other, yet-to-be-discovered, changes in synaptic function. Finally, the report by Carta et al expands our understanding of how excitatory synaptic transmission is fine-tuned at an important central synapse and underscores the fact that even well-trod ground (or synapses) continue to yield surprises that inform our understanding of the remarkable mechanistic diversity underlying synaptic plasticity in the CNS.  相似文献   

7.
Cellular prion protein (PrPC) appears to be involved in numerous physiological processes. We have recently shown a novel modulation of NMDA receptors by PrPC that results in neuroprotection via silencing of NMDA receptors containing NR2D subunits, whereas no effects on AMPA receptor function could be observed (Khosravani, et al. J Cell Biol 2008; 181:551). Here we show that PrP-null mice show a normal response to long-term depression stimuli requiring AMPA receptor activity, thus further supporting our previous findings of a selective action on NMDA receptors among ionotropic glutamate receptors.Key words: AMPA receptor, NMDA receptor, PrP, long term depression, LTDThe role of prion proteins in the pathophysiology of transmissible spongiform encephalopathies is well documented.1 Although there is a growing body of literature associating normal cellular prion protein (PrPC) with functions such as regulation of cell proliferation and survival, cell signalling and immune function,1 the spectrum of physiological roles attributable to PrPC remains to be determined. This may in part be due to the fact that mice lacking PrPC display a relatively mild phenotype, unless subjected to insults such as ischemia or seizures, where increased mortality of the PrP-null mice has been reported.25 Interestingly, the increased neuronal damage in PrP-null mice following excitotoxicity is alleviated upon treatment with the N-Methyl-D-Aspartate (NMDA) receptor (NMDAR) inhibitor MK-801,6 suggesting a neuroprotective role of PrPC via an action on NMDARs, but the mechanism was unclear.We recently described a novel action of PrPC on NMDAR function.7 By examining the neurophysiological properties of hippocampal neurons isolated from PrP-null mice, we were able to show that PrP-null mouse neurons exhibit enhanced and drastically prolonged NMDA evoked currents due to a functional upregulation of NMDARs containing NR2D subunits. Biochemical analyses suggested that NR2D subunits, but not NR2B subunits, co-immunoprecipitated with PrPC, indicating that PrPC and NMDARs form physical signaling complexes in neurons. The increased NMDAR function could be phenocopied by RNA interference and were rescued upon overexpression of exogenous PrPC. The enhanced NMDAR activity resulted in increased neuronal excitability, as well as enhanced glutamatergic-based excitotoxicity in both in vitro and in vivo experiments were neurons were transiently exposed to the selective agonist NMDA. Hence, native PrPC appears to mediate an important neuroprotective role by virtue of its ability to silence NR2D containing NMDARs. In contrast, minor effects on amplitude and rise and decay-time kinetics were observed for both α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) and GABAA (miniature and evoked) currents in synaptically mature hippocampal cultures.AMPA and NMDA receptors have been linked to synaptic plasticity, in particular long term potentiation (LTP) and long term depression (LTD). LTP is believed to mostly reflect a strengthening of the postsynaptic response, caused by a brief period of hyperexcitability that releases significant amounts of glutamate such as during a brief tetanic stimulation. This is thought to result in the opening of AMPA receptors, which depolarize the postsynaptic membrane. This in turn increases the activity of postsynaptic NMDARs, because magnesium ions that normally inhibit NMDAR activity, are dislodged by the postsynaptic depolarization, thus allowing NMDARs to become active. This functional activation of NMDARs results in the influx of calcium ions, which in turn initiate a signaling cascade that results in the membrane insertion of additional AMPA receptors, thus strengthening the synapse. This process is thought to involve NMDA receptor isoforms that predominantly contain the NR2A subunits.8A synaptic depotentiation process also can take place that results in the opposite effects of LTP; this process is known as long-term depression or LTD, which also has an NMDAR-dependent component. In contrast to brief tetanic stimulation, as is used in the induction of LTP, establishing LTD requires low frequency stimulation (e.g., 1 Hz for 15 min). Successful and repeatable induction of LTD depends on the parameters used for the conditioning stimulus and more importantly on the age of the animal. In juvenile animals (P12–P21) a low frequency protocol is effective and the mechanism of LTD is believed to depend on the activity of NMDA receptors containing NR2B subunits.8 Although a clear distinction of the roles between NR2A and NR2B containing NMDARs, in LTP and LTD respectively, has remained controversial,9 it is clear that both NR2A and NR2B are key mediators of alterations in synaptic plasticity. In older animals, the conditioning protocol is reported to require modification to include paired-pulses. This is thought to be due to the involvement of predominantly AMPA (and perhaps kainate) receptors in addition to mGluRs responsible for the synaptic depotentiation (reviewed in ref. 10).As mentioned earlier, our data obtained from hippocampal cultures indicated only a minor effect of PrPC knockout on AMPA receptor function. Hence, we hypothesized that AMPA receptor-mediated LTD should be similar in both wild type and PrP-null mice. We therefore examined the effect of PrP on LTD in hippocampal slices obtained from P30–P45 wild type mice and Zurich 1 PrPC knockout mice. Extracellular potentials were recorded using a patch pipette filled with 150 mM NaCl. First, 10 minutes of baseline evoked (every 30 sec) potentials were recorded to ensure stability of the preparation. LTD was then evoked by application of conditioning paired pulses (Δt = 60 ms) delivered at 1 Hz for 15 min. Thereafter, the field response was sampled every 30 sec for 40 min. As shown in Figure 1, this protocol evoked reliable LTD in wild type mice that partially recovered over the time course of about 20 minutes (Fig. 1). In the PrPC-null slices, LTD was indistinguishable from that observed in the wild type slices (Fig. 1).Open in a separate windowFigure 1LTD in the CA1 region of hippocampal slices from adult (P30–P45) WT and PrP-null mice. The conditioning pulse (arrow head) was delivered as paired-pulses (Δt = 60 ms) at 1 Hz for 15 min at the Schaffer collaterals. Analysis of field excitatory postsynaptic potential (fEPSP) slope revealed no statistically significant differences (Student''s t-test, p > 0.05) in the extent of induced LTD or the time course of its recovery to baseline. Numbers in parentheses indicate number of slices.The age of the animals, combined with the paired pulse protocol used in our experiments was designed to isolate AMPA receptor mediated LTD.10 The notion that LTD was unaltered in PrP-null mice is consistent with the observation that AMPA currents were not altered in these mice, and that AMPA receptor-mediated spontaneous synaptic events showed only minute changes compared with wild type animals. These data are also consistent with the notion that PrP-null mice show only mild phenotypes in spatial learning, with no apparent overall short-term memory deficits. Collectively, these data further support a selective action of PrP on NMDA receptors, rather than overall glutamatergic synaptic transmission.  相似文献   

8.
Metabotropic glutamate receptors type 1 (mGluR1s) are required for a normal function of the mammalian brain. They are particularly important for synaptic signaling and plasticity in the cerebellum. Unlike ionotropic glutamate receptors that mediate rapid synaptic transmission, mGluR1s produce in cerebellar Purkinje cells a complex postsynaptic response consisting of two distinct signal components, namely a local dendritic calcium signal and a slow excitatory postsynaptic potential. The basic mechanisms underlying these synaptic responses were clarified in recent years. First, the work of several groups established that the dendritic calcium signal results from IP3 receptor-mediated calcium release from internal stores. Second, it was recently found that mGluR1-mediated slow excitatory postsynaptic potentials are mediated by the transient receptor potential channel TRPC3. This surprising finding established TRPC3 as a novel postsynaptic channel for glutamatergic synaptic transmission.Glutamate is the predominant neurotransmitter used by excitatory synapses in the mammalian brain (Hayashi 1952; Curtis et al. 1959). At postsynaptic sites, glutamate binds to two different classes of receptors, namely the ionotropic glutamate receptors (iGluRs) and the metabotropic glutamate receptors (mGluRs) (Sladeczek et al. 1985; Nicoletti et al. 1986; Sugiyama et al. 1987). The iGluRs represent ligand-gated nonselective cation channels that underlie excitatory postsynaptic currents (EPSCs). Based on their subunit composition, gating, and permeability properties, they are subdivided into three groups named after specific agonists: AMPA- (α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid), NMDA receptors (N-methyl D-aspartate receptors) and kainate receptors (Alexander et al. 2009). The other class of glutamate receptors, the mGluRs, consists of receptors that are coupled to G proteins and act through distinct downstream signaling cascades. They are structurally different from iGluRs and characterized by the presence of seven transmembrane domains (Houamed et al. 1991; Masu et al. 1991). The mGluRs exist as homodimers that do not by themselves form an ion-permeable pore in the membrane (Ozawa et al. 1998). To date, eight different genes (and more splice variants) encoding mGluRs have been identified and form the mGluR1 through mGluR8 subtypes (Alexander et al. 2009). Based on the amino acid sequence homology, downstream signal transduction pathways, and pharmacological properties, each of the subtypes was assigned to one of three groups. Group I receptors consist of mGluR1 and mGluR5 that positively couple to the phospholipase C (PLC). The receptors mGluR2 and mGluR3 constitute group II, whereas the remaining mGluRs, namely mGluR4, mGluR6, mGluR7, and mGluR8, belong to group III. Both groups II and III inhibit the adenylyl cyclase and thereby reduce the concentration of cAMP in the cytosol.Of all different subtypes, mGluR1 is the most abundantly expressed mGluR in the mammalian central nervous system. In the brain, mGluR1 is highly expressed in the olfactory bulb, dentate gyrus, and cerebellum (Lein et al. 2007). The highest expression level of mGluR1 in the brain is found in Purkinje cells, the principal neurons of the cerebellar cortex (Shigemoto et al. 1992; Lein et al. 2007). Together with the AMPA receptors, mGluR1s are part of the excitatory synapses formed between parallel fibers and Purkinje cells (Fig. 1A). Each Purkinje cell is innervated by 100,000–200,000 parallel fibers (Ito 2006) that are axons of the cerebellar granule cells, the most abundant type of neuron in the brain. A second type of excitatory input to Purkinje cells is represented by the climbing fibers that originate in the inferior olive in the brain stem (Ito 2006). The two excitatory synaptic inputs to Purkinje cells are important determinants for the main functions of the cerebellum, including the real-time control of movement precision, error-correction, and control of posture as well as the procedural learning of complex movement sequences and conditioned responses.Open in a separate windowFigure 1.Parallel fiber-evoked mGluR1-dependent signals. (A) Diagram showing the parallel fiber synaptic input to Purkinje cell dendrites. (B) Microelectrode recording of glutamatergic postsynaptic potentials from a Purkinje cell in an acute slice of adult rat cerebellum. Short trains of stimuli to the parallel fibers (5–6 at 50 Hz) caused summation of the early AMPA receptor-dependent EPSPs (leading to spike firing) and a slow, delayed, depolarizing potential (slow EPSP), which was reversibly inhibited by antagonist of mGluRs (+)-MCPG (1mM). (C) Confocal image of a patch-clamped Purkinje cell in a cerebellar slice of an adult mouse. The patch-clamp pipette and the glass capillary used for electrical stimulation of parallel fibers are depicted schematically. The site of stimulation is shown at higher magnification in D. (D) Left: Parallel fiber-evoked (five pulses at 200 Hz, in 10 mM CNQX) synaptic responses consisting of a dendritic mGluR1-dependent Ca2+ transient (ΔF/F, top) and an early rapid and a slow excitatory postsynaptic current (EPSC, bottom). Block of the mGluR1-dependent components by the group I-specific mGluR-antagonist CPCCOEt (200 µM) is shown as indicated. Right: Pseudocolor image of the synaptic Ca2+ signal. (B, Reprinted with modifications, with permission, from Batchelor and Gaithwaite 1997 [Nature Publishing Group].)It is expected that mGluR1 is involved in many of these cerebellar functions. This view is supported by the observation that mGluR1-deficient knockout mice show severe impairments in motor coordination. In particular, the gait of these mice is strongly affected as well as their ability for motor learning and general coordination (Aiba et al. 1994). The phenotype of the general mGluR1-knockout mice is rescued by the insertion of the gene encoding mGluR1 exclusively into cerebellar Purkinje cells (Ichise et al. 2000) and blockade of mGluR1 expression only in Purkinje cells of adult mice leads to impaired motor coordination (Nakao et al. 2007). These findings established mGluR1 in Purkinje cell as synaptic receptors that are indispensable for a normal cerebellar function.Synaptic transmission involving mGluR1s is found at both parallel fiber-Purkinje cell synapses (Batchelor and Garthwaite 1993; Batchelor et al. 1994) as well as at climbing fiber-Purkinje cell synapses (Dzubay and Otis 2002). Most of our knowledge on the mGluR1 was gained from the analysis of the parallel fiber synapses. The parallel fiber synapse is quite unique in the central nervous system regarding its endowment with neurotransmitter receptors. In contrast to most other glutamatergic synapses in the mammalian brain, it lacks functional NMDA receptors (Shin and Linden 2005). The entire synaptic transmission at these synapses relies on AMPA receptors and on mGluR1 (Takechi et al. 1998). Although AMPA receptors are effectively activated even with single shock stimuli (Konnerth et al. 1990; Llano et al. 1991b), activation of mGluRs requires repetitive stimulation (Batchelor and Garthwaite 1993; Batchelor et al. 1994; Batchelor and Garthwaite 1997; Takechi et al. 1998). A possible explanation for the need of repetitive stimulation may relate to the observation that mGluR1s are found mostly at the periphery of the subsynaptic region (Nusser et al. 1994). At these sites outside the synaptic cleft, glutamate levels that are sufficiently high for receptor activation may be reached only with repetitive stimulation.At parallel fiber-Purkinje cell synapses, repetitive stimulation produces an initial AMPA receptor postsynaptic signal component, followed by a more prolonged mGluR1 component (Fig. 1). Figure 1B shows a current clamp recording of this response consisting of an early burst of action potentials, followed by a prolonged depolarization known as a “slow excitatory postsynaptic potential” (slow EPSP) (Batchelor and Garthwaite 1993; Batchelor et al. 1994; Batchelor and Garthwaite 1997). Voltage-clamp recordings allow a clear separation of the initial rapid, AMPA receptor mediated excitatory postsynaptic current (EPSC) and the mGluR1-mediated slow EPSC (Fig. 1D) (Takechi et al. 1998; Hartmann et al. 2008). In addition of inducing the slow EPSPs, mGluR1s mediate a large and highly localized dendritic calcium transient in cerebellar Purkinje cells (Fig. 1D) (Llano et al. 1991a; Finch and Augustine 1998; Takechi et al. 1998).  相似文献   

9.
10.
Signal transduction is regulated by protein–protein interactions. In the case of the ErbB family of receptor tyrosine kinases (RTKs), the precise nature of these interactions remains a topic of debate. In this review, we describe state-of-the-art imaging techniques that are providing new details into receptor dynamics, clustering, and interactions. We present the general principles of these techniques, their limitations, and the unique observations they provide about ErbB spatiotemporal organization.Signal transduction is associated with dramatic spatial and temporal changes in membrane protein distribution. Although the biochemical events downstream of membrane receptor activation are often well characterized, the initiating events within the plasma membrane remain unclear. Many cell surface receptors have been shown to redistribute into clusters in response to ligand binding (Metzger 1992). Therefore, correlating membrane receptor activation with dynamics and aggregation state is essential to understanding cell signaling.The role of receptor aggregation is of particular interest in the case of receptor tyrosine kinases (RTKs). It is generally accepted that ligand binding to the extracellular domain of RTKs induces dimerization, whether ligand- or receptor-mediated (Lemmon and Schlessinger 2010). However, there is evidence that some RTKs exist as oligomers in the absence of ligand, whereas others require higher-order oligomerization for activation (Lemmon and Schlessinger 2010). Understanding the fundamental interactions that regulate RTK signaling still remains an important focus in the field.Over the past decade, imaging technologies and biological tools have developed to a point such that questions about protein dynamics, clustering, and interactions can now be addressed in living cells (Fig. 1). These techniques reveal information about protein behavior on a spatial and temporal scale that is not provided by traditional biochemical assays. In this review, we will discuss the application of these advanced imaging technologies to the study of the ErbB family of RTKs.Open in a separate windowFigure 1.Summary of imaging techniques for quantifying receptor clustering, dynamics, and interactions.  相似文献   

11.
The discovery of the genetic causes of syndromic autism spectrum disorders and intellectual disabilities has greatly informed our understanding of the molecular pathways critical for normal synaptic function. The top-down approaches using human phenotypes and genetics helped identify causative genes and uncovered the broad spectrum of neuropsychiatric features that can result from various mutations in the same gene. Importantly, the human studies unveiled the exquisite sensitivity of cognitive function to precise levels of many diverse proteins. Bottom-up approaches applying molecular, biochemical, and neurophysiological studies to genetic models of these disorders revealed unsuspected pathogenic mechanisms and identified potential therapeutic targets. Moreover, studies in model organisms showed that symptoms of these devastating disorders can be reversed, which brings hope that affected individuals might benefit from interventions even after symptoms set in. Scientists predict that insights gained from studying these rare syndromic disorders will have an impact on the more common nonsyndromic autism and mild cognitive deficits.It is estimated that ∼1% of the human population has an autism spectrum disorder (ASD). ASD has widely varied behavioral manifestations, severity, and comorbid conditions (hence the term “spectrum”), but those diagnosed with autism are characterized by impaired communication and reciprocal social interactions, and restricted and repetitive patterns of activities and interests (Baird et al. 2006). Approximately 70% of those diagnosed with autism also have intellectual disability (ID), and 25% have a seizure disorder (Tuchman and Rapin 2002). There is a strong genetic basis for autism, but the risk architecture is highly heterogeneous, and a large number of genes have been implicated (Abrahams and Geschwind 2008). This daunting phenotypic and etiologic complexity, shared by other major psychiatric illnesses, has slowed progress toward developing new therapies.However, autism researchers are optimistic that the possibility of substantial progress may soon be realized (Krueger and Bear 2011). First, the genes have been discovered for numerous syndromic disorders that prominently feature ASD and ID. Second, these gene mutations have been reproduced in animal models that allow detailed examination of the underlying brain pathophysiology. Third, animal research has converged on altered synaptic function as a likely basis for impaired cognition and possibly ASD. Fourth, insights gained on how synapses function differently in the face of these mutations have suggested novel therapeutic interventions validated in preclinical models and that have shown promise in preliminary human clinical trials. Fifth, the fact that ASD and ID can be diagnosed in early childhood maximizes potential benefits of therapy because it can be started at a time when the brain is most plastic. Finally, animal studies using gene reactivation or pharmacological interventions suggest that substantial improvements can be seen even when treatments begin in adulthood (Ehninger et al. 2008b). Thus, a genetic diagnosis of a developmental brain disorder need not be a “life sentence” of permanent and inexorable mental disability.The path from gene discovery to novel treatment is outlined in Figure 1. This process often begins with astute clinical observations that some patients can be distinguished by a common set of phenotypic traits, thus defining a syndrome. Molecular genetic studies can then be undertaken to test the hypothesis that the syndrome has a genetic cause. In the event that disruption of a single gene or DNA segment causes the disease (i.e., a “highly penetrant” mutation), then it is possible to create an animal model (usually a mouse) that carries the same genetic disruption. Although the effects of the genetic lesion will likely manifest differently at the behavioral level in animals and humans because of differences in the complexity of the brains, it is reasonable to postulate that disruptions in elementary neuronal functions are likely to be shared. Understanding this neuronal pathophysiology is critical for identifying potential therapeutic targets. If these targets can be validated in the animal models, then chemistry ensues to generate molecules that can engage the target and satisfy the pharmacodynamic and pharmacokinetic drug requirements. If they are shown to be safe, drug candidates may then advance to human clinical trials. There are currently clinical trials ongoing in several single-gene syndromic disorders associated with ASD and ID. Most of these target alterations in synaptic signaling.Open in a separate windowFigure 1.The promise of molecular medicine in genetically defined disorders of brain development.The notion that some ASD and associated ID represent “synapsopathies” (or “synaptopathies”) is supported by the preponderance of penetrant mutations in genes associated with synaptic structure and function. The most common single-gene mutations in ASD with ID are associated with fragile X syndrome (FMR1), tuberous sclerosis (TSC1, TSC2), neurofibromatosis (NF1), Angelman syndrome (UBE3A), Rett syndrome (MECP2), the PTEN hamartoma tumor syndrome, and Phelan-McDermid syndrome (SHANK3) (for review, see Betancur 2011). Rare mutations in the neuroligin (NLGN3, NLGN2) and neurexin (NRXN1) genes also cause autism (Jamain et al. 2003). Although this is by no means an exhaustive list of genes implicated in autism (and many await discovery), it is notable nonetheless that these highly penetrant mutations occur in genes that are critical regulators of synaptic function, and further, illuminate biochemical pathways that might be pathogenic in ASD and ID (Fig. 2).Open in a separate windowFigure 2.(A) Schematic of a neuron and axonal-dendritic synapse that depict examples of cellular localization of the various types of defects in ASD/ID. (B) A signaling pathway at the excitatory synapses that couples activity as registered by the release of glutamate to local control of protein synthesis. Disruption of the gene products indicated in the colored boxes greatly increases the risk of ASD/ID. Syndromic disorders with increased prevalence of ASD include Phelan-McDermid Syndrome (SHANK3); Noonan syndrome (RAF1, MEK1); Neurofibromatosis type 1 (NF1); Costello syndrome (H-Ras, MEK1); Cowden syndrome (PTEN); Cardio-facio-cutaneous (CFC) syndrome (MEK1/2); Tuberous sclerosis complex (TSC1/2); Fragile X syndrome (FMRP); Angelman syndrome (AS UBE3a); Rett syndrome (RTT–MeCP2); and Rubinstein-Taybi syndrome (RTS–CREB binding protein, p300). Rare, nonsyndromic ASDs include NLGN3/4 and NRXN1; ID/ASD: SHANK2.In this article, we focus on a few syndromic disorders associated with ASD and ID that are characterized by penetrant mutations in genes that have been shown in animal models to disrupt synaptic function. Our goal is to highlight the similarities and differences in these syndromes and their underlying synaptic pathophysiology. Optimal synaptic function occurs within a narrow dynamic range along many dimensions, and it is not surprising that pathophysiology occurs at the edges of these spectra. What has come as a surprise, however, is that ASD and ID appear to be common consequences of disruptive mutations that cause synaptic pathophysiology at both ends of a spectrum. In other words, both “gain-of-function” and “loss-of-function” mutations can manifest in similar ways. Insights into the pathophysiology of ASD and ID have raised the possibility of therapeutic interventions to bring synapses into a normal operating range.  相似文献   

12.
The establishment of precise neuronal cell morphology provides the foundation for all aspects of neurobiology. During development, axons emerge from cell bodies after an initial polarization stage, elongate, and navigate towards target regions guided by a range of environmental cues. The Rho and Ras families of small GTPases have emerged as critical players at all stages of axonogenesis. Their ability to coordinately direct multiple signal transduction pathways with precise spatial control drives many of the activities that underlie this morphogenetic program: the dynamic assembly, disassembly, and reorganization of the actin and microtubule cytoskeletons, the interaction of the growing axon with other cells and extracellular matrix, the delivery of lipids and proteins to the axon through the exocytic machinery, and the internalization of membrane and proteins at the leading edge of the growth cone through endocytosis. This article highlights the contribution of Rho and Ras GTPases to axonogenesis.The Ras superfamily of small GTPases, consisting of almost 200 proteins, can be subclassified into six families: Rho, Ras, Rab, Arf, Sar, and Ran (Colicelli 2004). These proteins act as molecular switches, cycling between an inactive, GDP-bound state and an active, GTP-bound state (Fig. 1). The activated conformation interacts with specific effectors to propagate downstream signaling events that influence many aspects of cell biology. Guanine nucleotide exchange factors (GEFs) activate the switch by catalyzing the exchange of GDP for GTP, whereas GTPase-activating proteins (GAPs) increase the intrinsic GTPase activity and inactivate the switch (Fig. 1) (Jaffe and Hall 2005). Dominant–negative (DN) and constitutively active (CA) versions of small GTPases (created through specific amino acid substitutions) have been used extensively to dissect the individual roles of these proteins. Although these have been incredibly informative, they do have potential drawbacks: dominant–negative constructs, which act by sequestering GEFs, may interfere with closely related family members, whereas constitutively activated GTPases interact indiscriminately with all their potential targets, something that does not happen under normal conditions. RNAi and gene knockout approaches afford the potential for greater specificity, but they too have limitations, because GTPases, their regulators, and their targets are typically found as closely related isoforms. This article focuses on the role of Rho and Ras family members in four different aspects of axonogenesis: initiation, elongation, guidance, and branching. The major role of Rho GTPases, conserved in all eukaryotes, is to control the assembly, disassembly, and dynamic rearrangements of the actin and microtubule cytoskeletons. It is not surprising, therefore, that they play crucial roles in the growth, guidance, and branching of axons. Ras GTPases, on the other hand, are activated by a large number of plasma membrane growth factor receptors and adhesion receptors to promote key signal transduction pathways, including ERK, MAP kinase, and PI3-kinase, which play a variety of important roles in axonogenesis.Open in a separate windowFigure 1.The GTPase cycle GTPases. (Ras, in this example) cycle between an inactive GDP-bound state and an active, GTP-bound state. Following a specific stimulus, GEFs catalyze the exchange of GDP for GTP, enabling the interaction of GTPases with specific effectors leading to cellular responses. In contrast, GAPs inactivate GTPases by stimulating their intrinsic GTPase activity.  相似文献   

13.
Long-term potentiation (LTP) and long-term depression (LTD) are the major forms of functional synaptic plasticity observed at CA1 synapses of the hippocampus. The balance between LTP and LTD or "metaplasticity" is controlled by G-protein coupled receptors (GPCRs) whose signal pathways target the N-methyl-D-asparate (NMDA) subtype of excitatory glutamate receptor. We discuss the protein kinase signal cascades stimulated by Galphaq and Galphas coupled GPCRs and describe how control of NMDAR activity shifts the threshold for the induction of LTP.  相似文献   

14.
Chemotaxis—the directed movement of cells in a gradient of chemoattractant—is essential for neutrophils to crawl to sites of inflammation and infection and for Dictyostelium discoideum (D. discoideum) to aggregate during morphogenesis. Chemoattractant-induced activation of spatially localized cellular signals causes cells to polarize and move toward the highest concentration of the chemoattractant. Extensive studies have been devoted to achieving a better understanding of the mechanism(s) used by a neutrophil to choose its direction of polarity and to crawl effectively in response to chemoattractant gradients. Recent technological advances are beginning to reveal many fascinating details of the intracellular signaling components that spatially direct the cytoskeleton of neutrophils and D. discoideum and the complementary mechanisms that make the cell''s front distinct from its back.Chemotaxis—the directed movement of cells in a gradient of chemoattractant—allows leukocytes to seek out sites of inflammation and infection, amoebas of Dictyostelium discoideum (D. discoideum) to aggregate, neurons to send projections to specific regions of the brain to find their synaptic partners, yeast cells to mate, and fibroblasts to move into the wound space (Fig. 1). In each case, chemoattractant-induced activation of spatially localized cellular signals causes cells to polarize and move toward the highest concentration of the chemoattractant. During chemotaxis, filamentous actin (F-actin) is polymerized asymmetrically at the upgradient edge of the cell (leading edge), providing the necessary force to thrust projections of the plasma membrane in the proper direction (see Mullins 2009). Neutrophilic leukocytes (neutrophils), for instance, can polarize and move up very shallow gradients, with a chemoattractant concentration ∼2% higher at the front than the back (Fig. 2) (Devreotes and Zigmond 1988). To restrict actin polymerization to the leading edge in such a shallow gradient, neutrophils must create a much steeper internal gradient of regulatory signals. In addition, distinctive actin–myosin contractile complexes are also formed at the sides and back of the cells (Fig. 2). The ability to create such distinctive segregation of actin assemblies enables neutrophils to move nearly 50 times more quickly than fibroblasts. The polarization response is self-organizing, which occurs even when the attractant concentration is uniform and apparently stimulating all portions of the plasma membrane at the same intensity; in the absence of a gradient, the direction of polarity is random, but all cells can be induced to polarize (Fig. 2). Thus, neutrophil polarization to chemoattractant stimulation represents a striking example of symmetry breaking from an unpolarized state to a polarized one.Open in a separate windowFigure 1.Examples of chemotaxis. (A) A human neutrophil chasing a Staphylococcus aureus microorganism on a blood film among red blood cells, notable for their dark color and principally spherical shape (imaged by David Rogers, courtesy of Thomas P. Stossel). Bar, 10 µm. Chemotaxis is also necessary for (B) D. discoideum to form multicellular aggregates during development (courtesy of M.J. Grimson and R.L. Blanton, Texas Tech University), and (C) for axons to find their way in the developing nervous system. Photo provided by Kathryn Tosney, University of Miami.Open in a separate windowFigure 2.(A–D) Polarization of a neutrophil in response to gradient of chemoattractant. Nomarski images of unpolarized neutrophil responding to a micropipette containing the chemoattractant fMLP (white circle) at (A) 5 s, (B) 30 s, (C) 81 s, and (D) 129 s of stimulation. Bar = 5 µm. (Figure is taken from Weiner et al. 1999, with permission.) Human neutrophils stimulated with fMLP show highly polarized morphology and asymmetric cytoskeletal assemblies. (E–G) Human neutrophils were stimulated by a uniform concentration of fMLP (100 nM) and fixed 2 min after stimulations. Fixed cells were stained for F-actin with rhodamine-phalloidin (E, red) and an antibody raised against activated myosin II (phosphorylated specifically at Ser19, p[19]-MLC) (F, green). These fluorescent images are merged with Nomarski image in (G). Bars, 10 µm.To enter an infected tissue, neutrophils require chemoattractants produced by host cells and microorganisms to migrate to the sites and infection and inflammation. Neutrophil chemotaxis also contributes to many inflammatory and autoimmune diseases, including rheumatoid arthritis, ischemia-reperfusion syndrome, acute respiratory distress, and systemic inflammatory response syndromes. Although the critical physiological functions of neutrophils have made their chemoattractants and chemoattractant receptors targets of intense investigation, understanding of the neutrophil polarity and directional migration has until recently lagged behind that of other cells. Over the past decade, experimentation with knockout mice and human neutrophil cell lines has begun to shed light on the complex intracellular signals responsible for neutrophil polarity. In this article, I summarize recent advances in the study of chemotactic signals in neutrophils, with some of the discussion also devoted to a related model—chemotaxis of D. discoideum. These soil amoebas grow as single cells, but on starvation chemotax into multicellular aggregates in response to secreted chemoattractants such as adenosine 3′,5′-monophosphate (cAMP).  相似文献   

15.
How are the asymmetric distributions of proteins, lipids, and RNAs established and maintained in various cell types? Studies from diverse organisms show that Par proteins, GTPases, kinases, and phosphoinositides participate in conserved signaling pathways to establish and maintain cell polarity.The asymmetric distribution of proteins, lipids, and RNAs is necessary for cell fate determination, differentiation, and specialized cell functions that underlie morphogenesis (St Johnston 2005; Gonczy 2008; Knoblich 2008; Macara and Mili 2008; Martin-Belmonte and Mostov 2008). A fundamental question is how this asymmetric distribution is established and maintained in different types of cells and tissues. The formation of a specialized apical surface on an epithelial cell seems quite different from the specification of axons versus dendrites in a neuron, or the asymmetric division of a nematode zygote. Yet, remarkably, a conserved molecular toolbox is used throughout the metazoa to establish and maintain cell polarity in these and many other contexts. This toolbox consists of proteins that are components of signal transduction pathways (Goldstein and Macara 2007; Assemat et al. 2008; Yamanaka and Ohno 2008). However, our understanding of these pathways, and their intersection with other signaling networks, remains incomplete. Moreover, the regulation and cross talk between the polarity proteins and other signaling components varies from one context to another, which complicates the task of dissecting polarity protein function. Nonetheless, rapid progress is being made in our understanding of polarity signaling, which is outlined in this article, with an emphasis on the Par proteins, because these proteins play major roles integrating diverse signals that regulate cell polarity (Fig. 1) (see Munro and Bowerman 2009; Prehoda 2009; Nelson 2009).Open in a separate windowFigure 1.An overview of Par complex signaling, showing inputs (bottom) and outputs (top) with cellular functions that are targeted by these pathways (italics).  相似文献   

16.
The mammary gland is an organ that at once gives life to the young, but at the same time poses one of the greatest threats to the mother. Understanding how the tissue develops and functions is of pressing importance in determining how its control mechanisms break down in breast cancer. Here we argue that the interactions between mammary epithelial cells and their extracellular matrix (ECM) are crucial in the development and function of the tissue. Current strategies for treating breast cancer take advantage of our knowledge of the endocrine regulation of breast development, and the emerging role of stromal–epithelial interactions (Fig. 1). Focusing, in addition, on the microenvironmental influences that arise from cell–matrix interactions will open new opportunities for therapeutic intervention. We suggest that ultimately a three-pronged approach targeting endocrine, growth factor, and cell-matrix interactions will provide the best chance of curing the disease.Cellular interactions with the ECM are one of the defining features of metazoans (Huxley-Jones et al. 2007). Matrix proteins are among the most abundant in the body, and are integral components of cell regulation and developmental programs operating in all tissues. They provide structure and support to tissues, and they interact with cells through diverse receptors to guide development, patterning, and cell fate decisions (Streuli 2009). Together with cytokines and growth factors, and cell–cell interactions, the ECM determines whether cells survive, proliferate, differentiate, or migrate, and it influences cell shape and polarity (Streuli and Akhtar 2009). Cell–ECM interactions also are central in the assembly of the matrix itself, and in determining ECM organization and rigidity (Kadler et al. 2008; Kass et al. 2007). The cell–matrix interface is therefore pivotal in controlling both cell function and tissue structure, which together build organs into operational structures. Thus, elucidating precisely how the matrix directs cell phenotype is crucial for understanding mechanisms of development and disease.Mammary gland tissue contains epithelium and stroma ((Fig.Fig. 2). Mammary epithelial cells (MEC) form collecting ducts and, in pregnancy and lactation, milk-secreting alveoli (or lobules). The mammary epithelium is bilayered, with the inner luminal cells facing a central apical cavity and surrounded by the outer basal, myoepithelial cells. It also harbors stem and progenitor cells, which are the source of both luminal and myoepithelial cells (Visvader 2009). The epithelium is ensheathed by one of the main types of ECM, basement membrane (BM), which separates epithelium from stroma, and profoundly influences the development and biology of the gland (Streuli 2003). The stroma includes fibrous connective tissue ECM proteins, and a wide variety of cell types, including inter- and intralobular fibroblasts, adipocytes, endothelial cells, and innate immune cells (both macrophages and mast cells). The stroma is the support network for the epithelium, providing both nutrients and blood supply, and immune defenses, as well as physical structure to the gland. Importantly, each of the different stromal cell types secrete instructive signals that are crucial for various aspects of the development and function of the epithelium (Sternlicht 2006).Open in a separate windowFigure 1.Mammary gland development. Whole mounts of (A) virgin and (B) mid-pregnant mouse mammary gland. The thin, branched epithelial ducts that are characteristic of nonpregnant gland undergo dramatic alterations in pregnancy, when new types of epithelial structures, the milk-producing alveoli, emerge. The huge amount of proliferation that accompanies this change occurs in a discrete and controlled fashion. The formation of ducts and alveoli is under three types of environmental control. The first is long-range endocrine hormones, which includes estrogen, progesterone, glucocorticoids, and prolactin. The second is locally acting growth factors, which arise from stromal–epithelial conversation, and includes amphiregulin, FGF, HGF, and IGF. Finally, microenvironmental adhesive signals from adjacent cells (e.g., via cadherins) and from the ECM (e.g., integrin) have an equally central role in all aspects of mammary development and function. Importantly, the proliferation that occurs in breast cancer is not well controlled, indicating not only defects in growth signaling, but also in cellular organization. Chronologically, breast cancer drugs were initially developed against endocrine regulators, e.g., estrogen, and more recently against the stromal/epithelial regulators, e.g., receptor tyrosine kinases. A complete control of the disease will only happen when therapies targeting the microenvironmental adhesion breast regulators, e.g., cell–matrix interactions, are formulated, and used in combination.Open in a separate windowFigure 2.Ducts and alveoli in early pregnancy. Transverse section of ducts surrounded by a thick layer of collagenous (stromal) connective tissue containing fibroblasts and the fat pad. Also visible are small alveoli, which fill the fat pad by the time the gland lactates, but note that they are not surrounded collagen. A capillary is evident, and macrophages and mast cells are also present, though they require specific staining to visualize. A basement membrane is present directly at the basal surface of both ductal and alveolar epithelium (see Fig. 3).BMs surround three cell types in the mammary gland: the epithelium, the endothelium of the vasculature, and adipocytes (Fig. 3). These ECMs are thin, ∼100-nm thick sheets of glycoproteins and proteoglycans, which are constructed around an assembled polymer of laminins and a cross-linked network of collagen IV fibrils (Yurchenco and Patton 2009). Laminins form αβγ trimers, and in the breast at least four distinct isoforms are present: laminin-111, -322, and -511 and -521 (previously known as LM-1, 5, 10, and 11) (Aumailley et al. 2005; Prince et al. 2002). Similarly, BM proteoglycans are diverse and show complexity in their GAG chain modifications that vary with development of the mammary gland, though the major species is perlecan (Delehedde et al. 2001). BM proteins interact with MEC via integrins and transmembrane proteoglycans dystroglycan and syndecan, which all couple to the cytoskeleton and assemble signaling platforms to control cell fate (Barresi and Campbell 2006; Morgan et al. 2007). The best-studied MEC BM receptors are integrins, which are αβ heterodimers: they include receptors for collagen (α1β1 and α2β1), LM-111, -511, -521 (α3β1, α6β1, and α6β4), LM-322 (α3β1 and α6β4), and in some MECs fibronectin and vitronectin (α5β1 and β3 integrins) (Naylor and Streuli 2006). BM proteoglycans have a further signaling role via their capacity to bind growth factors and cytokines: They act both as a reservoir and a delivery vehicle to GF receptors, thereby controlling the passage of GFs across the BM (Iozzo 2005). Because of these diverse roles, the BM is a dominant regulator of the mammary epithelial phenotype.Open in a separate windowFigure 3.Alveolar and ductal architecture of breast epithelia shown through fluorescence and histological images. (A) An alveolus from a lactating mammary gland, showing luminal epithelial cells with cell–cell adhesion junctions (green, E-cadherin) and cell–matrix interactions (red, laminin-111). The central lumen is where milk collects. (B) The duct of a nonpregnant gland is stained with an antibody to laminin (brown) and counterstained with hematoxylin. Note that the laminin-containing basement membrane surrounds the ductal epithelial cells, and outside this lie collagenous connective tissue and adipocytes. Figure B courtesy of Dr. Rama Khokha.Apart from the endothelium and adipocytes, which contact BMs, the mammary stromal cells are mostly solitary and embedded within a fibrous ECM. Stromal matrix components include collagens type I and III, proteoglycans and hyaluronic acid, fibronectin and tenascins, and the composition varies with development and pregnancy (Schedin et al. 2004). Not a great deal is known about the specific interactions between breast stromal cells and their ECM, or how the matrix composition and density determines stromal cell function. However, it is becoming evident that the stromal matrix exerts a powerful influence on malignant breast epithelial cells, which invade the stroma and are further transformed by exposure to this distinct microenvironment (Kumar and Weaver 2009; Streuli 2006).In this article we focus on cell–matrix interactions within mammary epithelium, and reveal known and possible mechanisms for its control on ductal development, alveolar function, and cancer progression.  相似文献   

17.
Auxin and Monocot Development   总被引:2,自引:0,他引:2  
Monocots are known to respond differently to auxinic herbicides; hence, certain herbicides kill broadleaf (i.e., dicot) weeds while leaving lawns (i.e., monocot grasses) intact. In addition, the characters that distinguish monocots from dicots involve structures whose development is controlled by auxin. However, the molecular mechanisms controlling auxin biosynthesis, homeostasis, transport, and signal transduction appear, so far, to be conserved between monocots and dicots, although there are differences in gene copy number and expression leading to diversification in function. This article provides an update on the conservation and diversification of the roles of genes controlling auxin biosynthesis, transport, and signal transduction in root, shoot, and reproductive development in rice and maize.Auxinic herbicides have been used for decades to control dicot weeds in domestic lawns (Fig. 1A), commercial golf courses, and acres of corn, wheat, and barley, yet it is not understand how auxinic herbicides selectively kill dicots and spare monocots (Grossmann 2000; Kelley and Reichers 2007). Monocots, in particular grasses, must perceive or respond differently to exogenous synthetic auxin than dicots. It has been proposed that this selectivity is because of either limited translocation or rapid degradation of exogenous auxin (Gauvrit and Gaillardon 1991; Monaco et al. 2002), altered vascular anatomy (Monaco et al. 2002), or altered perception of auxin in monocots (Kelley and Reichers 2007). To explain these differences, there is a need to further understand the molecular basis of auxin metabolism, transport, and signaling in monocots.Open in a separate windowFigure 1.Differences between monocots and dicots. (A) A dicot weed in a lawn of grasses. Note the difference in morphology of the leaves. (B) Germinating dicot (bean) seedling. Dicots have two cotyledons (cot). Reticulate venation is apparent in the leaves. The stem below the cotyledons is called the hypocotyl (hyp). (C) Germinating monocot (maize) seedling. Monocots have a single cotyledon called the coleoptile (col) in grasses. Parallel venation is apparent in the leaves. The stem below the coleoptile is called the mesocotyl (mes).Auxin, as we have seen in previous articles, plays a major role in vegetative, reproductive, and root development in the model dicot, Arabidopsis. However, monocots have a very different anatomy from dicots (Raven et al. 2005). Many of the characters that distinguish monocots and dicots involve structures whose development is controlled by auxin: (1) As the name implies, monocots have single cotyledons, whereas dicots have two cotyledons (Fig. 1B,C). Auxin transport during embryogenesis may play a role in this difference as cotyledon number defects are often seen in auxin transport mutants (reviewed in Chandler 2008). (2) The vasculature in leaves of dicots is reticulate, whereas the vasculature in monocots is parallel (Fig. 1). Auxin functions in vascular development because many mutants defective in auxin transport, biosynthesis, or signaling have vasculature defects (Scarpella and Meijer 2004). (3) Dicots often produce a primary tap root that produces lateral roots, whereas, in monocots, especially grasses, shoot-borne adventitious roots are the most prominent component of the root system leading to the characteristic fibrous root system (Fig. 2). Auxin induces lateral-root formation in dicots and adventitious root formation in grasses (Hochholdinger and Zimmermann 2008).Open in a separate windowFigure 2.The root system in monocots. (A) Maize seedling showing the primary root (1yR), which has many lateral roots (LR). The seminal roots (SR) are a type of adventitious root produced during embryonic development. Crown roots (CR) are produced from stem tissue. (B) The base of a maize plant showing prop roots (PR), which are adventitious roots produced from basal nodes of the stem later in development.It is not yet clear if auxin controls the differences in morphology seen in dicots versus monocots. However, both conservation and diversification of mechanisms of auxin biosynthesis, homeostasis, transport, and signal transduction have been discovered so far. This article highlights the similarities and the differences in the role of auxin in monocots compared with dicots. First, the genes in each of the pathways are introduced (Part I, Table I) and then the function of these genes in development is discussed with examples from the monocot grasses, maize, and rice (Part II).  相似文献   

18.
Many adult stem cells divide asymmetrically to balance self-renewal and differentiation, thereby maintaining tissue homeostasis. Asymmetric stem cell divisions depend on asymmetric cell architecture (i.e., cell polarity) within the cell and/or the cellular environment. In particular, as residents of the tissues they sustain, stem cells are inevitably placed in the context of the tissue architecture. Indeed, many stem cells are polarized within their microenvironment, or the stem cell niche, and their asymmetric division relies on their relationship with the microenvironment. Here, we review asymmetric stem cell divisions in the context of the stem cell niche with a focus on Drosophila germ line stem cells, where the nature of niche-dependent asymmetric stem cell division is well characterized.Asymmetric cell division allows stem cells to self-renew and produce another cell that undergoes differentiation, thus providing a simple method for tissue homeostasis. Stem cell self-renewal refers to the daughter(s) of stem cell division maintaining all stem cell characteristics, including proliferation capacity, maintenance of the undifferentiated state, and the capability to produce daughter cells that undergo differentiation. A failure to maintain the correct stem cell number has been speculated to lead to tumorigenesis/tissue hyperplasia via stem cell hyperproliferation or tissue degeneration/aging via a reduction in stem cell number or activity (Morrison and Kimble 2006; Rando 2006). This necessity changes during development. The stem cell pool requires expansion earlier in development, whereas maintenance is needed later to sustain tissue homeostasis.There are two major mechanisms to sustain a fixed number of adult stem cells: stem cell niche and asymmetric stem cell division, which are not mutually exclusive. Stem cell niche is a microenvironment in which stem cells reside, and provides essential signals required for stem cell identity (Fig. 1A). Physical limitation of niche “space” can therefore define stem cell number within a tissue. Within such a niche, many stem cells divide asymmetrically, giving rise to one stem cell and one differentiating cell, by placing one daughter inside and another outside of the niche, respectively (Fig. 1A). Nevertheless, some stem cells divide asymmetrically, apparently without the niche. For example, in Drosophila neuroblasts, cell-intrinsic fate determinants are polarized within a dividing cell, and subsequent partitioning of such fate determinants into daughter cells in an asymmetric manner results in asymmetric stem cell division (Fig. 1B) (see Fig. 3A and Prehoda 2009).Open in a separate windowFigure 1.Mechanisms of asymmetric stem cell division. (A) Asymmetric stem cell division by extrinsic fate determinants (i.e., the stem cell niche). The two daughters of stem cell division will be placed in distinct cellular environments either inside or outside the stem cell niche, leading to asymmetric fate choice. (B) Asymmetric stem cell division by intrinsic fate determinants. Fate determinants are polarized in the dividing stem cells, which are subsequently partitioned into two daughter cells unequally, thus making the division asymmetrical. Self-renewing (red line) and/or differentiation promoting (green line) factors may be involved.In this review, we focus primarily on asymmetric stem cell divisions in the Drosophila germ line as the most intensively studied examples of niche-dependent asymmetric stem cell division. We also discuss some examples of stem cell division outside Drosophila, where stem cells are known to divide asymmetrically or in a niche-dependent manner.  相似文献   

19.
Toll-like receptors sense pathogen-associated molecular patterns (e.g., lipopolysaccharides) and trigger gene-expression changes that ultimately eradicate the invading microbes.Toll-like receptors (TLRs) are protective immune sentries that sense pathogen-associated molecular patterns (PAMPs) such as unmethylated double-stranded DNA (CpG), single-stranded RNA (ssRNA), lipoproteins, lipopolysaccharide (LPS), and flagellin. In innate immune myeloid cells, TLRs induce the secretion of inflammatory cytokines (Newton and Dixit 2012), thereby engaging lymphocytes to mount an adaptive, antigen-specific immune response (see Fig. 1) that ultimately eradicates the invading microbes (Kawai and Akira 2010).Open in a separate windowFigure 1.TLR signaling (simplified view).Identification of TLR innate immune function began with the discovery that Drosophila mutants in the Toll gene are highly susceptible to fungal infection (Lemaitre et al. 1996). This was soon followed by identification of a human Toll homolog, now known as TLR4 (Medzhitov et al. 1997). To date, 10 TLR family members have been identified in humans, and at least 13 are present in mice. All TLRs consist of an amino-terminal domain, characterized by multiple leucine-rich repeats, and a carboxy-terminal TIR domain that interacts with TIR-containing adaptors. Nucleic acid–sensing TLRs (TLR3, TLR7, TLR8, and TLR9) are localized within endosomal compartments, whereas the other TLRs reside at the plasma membrane (Blasius and Beutler 2010; McGettrick and O’Neill 2010). Trafficking of most TLRs from the endoplasmic reticulum (ER) to either the plasma membrane or endolysosomes is orchestrated by ER-resident proteins such as UNC93B (for TLR3, TLR7, TLR8, and TLR9) and PRAT4A (for TLR1, TLR2, TLR4, TLR7, and TLR9) (Blasius and Beutler 2010). Once in the endolysosomes, TLR3, TLR7, and TLR9 are subject to stepwise proteolytic cleavage, which is required for ligand binding and signaling (Barton and Kagan 2009). For some TLRs, ligand binding is facilitated by coreceptors, including CD14 and MD2.Following ligand engagement, the cytoplasmic TIR domains of the TLRs recruit the signaling adaptors MyD88, TIRAP, TRAM, and/or TRIF (see Fig. 2). Depending on the nature of the adaptor that is used, various kinases (IRAK4, IRAK1, IRAK2, TBK1, and IKKε) and ubiquitin ligases (TRAF6 and pellino 1) are recruited and activated, culminating in the engagement of the NF-κB, type I interferon, p38 MAP kinase (MAPK), and JNK MAPK pathways (Kawai and Akira 2010; Morrison 2012). TRAF6 is modified by K63-linked autoubiquitylation, which enables the recruitment of IκB kinase (IKK) through a ubiquitin-binding domain of the IKKγ (also known as NEMO) subunit. In addition, a ubiquitin-binding domain of TAB2 recognizes ubiquitylated TRAF6, causing activation of the associated TAK1 kinase, which then phosphorylates the IKKβ subunit. Pellino 1 can modify IRAK1 with K63-linked ubiquitin, allowing IRAK1 to recruit IKK directly. TLR4 signaling via the TRIF adaptor protein leads to K63-linked polyubiquitylation of TRAF3, thereby promoting the type I interferon response via interferon regulatory factor (IRFs) (Hacker et al. 2011). Alternatively, TLR4 signaling via MyD88 leads to the activation of TRAF6, which modifies cIAP1 or cIAP2 with K63-linked polyubiquitin (Hacker et al. 2011). The cIAPs are thereby activated to modify TRAF3 with K48-linked polyubiquitin, causing its proteasomal degradation. This allows a TRAF6–TAK1 complex to activate the p38 MAPK pathway and promote inflammatory cytokine production (Hacker et al. 2011). TLR signaling is turned off by various negative regulators: IRAK-M and MyD88 short (MyD88s), which antagonize IRAK1 activation; FADD, which antagonizes MyD88 or IRAKs; SHP1 and SHP2, which dephosphorylate IRAK1 and TBK1, respectively; and A20, which deubiquitylates TRAF6 and IKK (Flannery and Bowie 2010; Kawai and Akira 2010).Open in a separate windowFigure 2.TLR signaling. (Adapted with kind permission of Cell Signaling Technology [http://www.cellsignal.com].)Deregulation of the TLR signaling cascade causes several human diseases. Patients with inherited deficiencies of MyD88, IRAK4, UNC93B1, or TLR3 are susceptible to recurrent bacterial or viral infections (Casanova et al. 2011). Chronic TLR7 and/or TLR9 activation in autoreactive B cells, in contrast, underlies systemic autoimmune diseases (Green and Marshak-Rothstein 2011). Furthermore, oncogenic activating mutations of MyD88 occur frequently in the activated B-cell-like subtype of diffuse large B-cell lymphoma and in other B-cell malignancies (Ngo et al. 2011). Inhibitors of various TLRs or their associated kinases are currently being developed for autoimmune or inflammatory diseases and also hold promise for the treatment of B-cell malignancies with oncogenic MyD88 mutations. Many TLR7 and TLR9 agonists are currently in clinical trials as adjuvants to boost host antitumor responses in cancer patients (Hennessy et al. 2010).  相似文献   

20.
The specificity of synaptic connections is directly related to the functional integrity of neural circuits. Long-range axon guidance and topographic mapping mechanisms bring axons into spatial proximity of target cells and thus limit the number of potential synaptic partners. Synaptic specificity is then achieved by extracellular short-range guidance cues and cell-surface recognition cues. Neural activity may enhance the precision and strength of specific circuit connections. Here, we focus on one of the final steps of synaptic matchmaking: the targeting of synaptic layers and the mutual recognition of axons and dendrites within these layers.Perception and behavior are critically dependent on synaptic communication between specific neurons. Understanding how neurons achieve such “synaptic specificity” is therefore one of the most fundamental issues in developmental neuroscience. Langley’s notion of “chemical relations” between synaptically connected neurons (Langley 1892) and Sperry’s “chemoaffinity” hypothesis (Sperry 1963) provided a conceptual framework for the development of precise synaptic connections in the central nervous system. Sperry postulated that molecular interactions between neurons and their extracellular environment (including between and amongst axons and dendrites) ensure that connections form only between “appropriate” synaptic partners (Sperry 1963). This hypothesis has been confirmed by experimental work over the last four decades, most importantly by the identification of molecular cues that provide synaptic specificity (see Sanes and Yamagata 2009 for a recent comprehensive review). However, within this broad framework, a number of alternate mechanisms have been shown or proposed to play roles in specific aspects of such targeting processes. Here, we focus on mechanisms that underlie the formation of synaptic layers, a prominent anatomical feature of the visual system as well as many other areas of the CNS.As reviewed previously (O''Leary 2010), the chemoaffinity principle underlies the developmental process of topographic mapping. Indeed, the precision with which neurons preserve the spatial relationships between the visual world and its representation in the brain is remarkable: Across animals ranging from flies to vertebrates, axons that bear signals from adjacent points in visual space invariably choose adjacent targets in the brain (Braitenberg 1967; Lemke and Reber 2005; Sperry 1963). Thus, position-dependent guidance of axons ensures that a visuotopic map develops. However, position in space is just one attribute of a visual stimulus; others include color, brightness, edge detection, and movement. If position in visual space is encoded by localized activation within a two-dimensional field of neurons, then these other features are encoded by local circuits that act both in series and in parallel and are reiterated many times across the field (Fig. 1). These local circuit modules are often envisioned as “columns” that lie orthogonal to the topographic map, with each column corresponding to a pixel in visual space and each level of the column representing a different, specific visual feature within that pixel, such as brightness, color, etc. (Fig. 1). How these columns acquire their laminated structure represents a developmental challenge of extraordinary scale. Although long-range axon guidance and topographic mapping no doubt contribute to restricting the astronomical number of potential synaptic partners, these mechanisms are clearly not sufficient; additional mechanisms must (and do) exist that act on a local scale to provide an additional level of positional information and cell-type-specific “chemoaffinity.”Open in a separate windowFigure 1.Laminae are a fundamental organizing unit of neural circuits. Each column corresponds to a single topographic position (e.g., location on the retina). Within each column, different cell types (shown type A: blue, and type B: red) respond to different features in the visual world, such as motion or luminance. These pixels are repeated many times over and thus cover all of visual space. A simple rule of “Cell type A connects to Layer A, etc.” ensures that functional segregation is maintained in the connections from the retina to the target (parallel processing). Each pixel P1, P2, and P3 connects to a single column (C1, C2, and C3), establishing serial processing. Within each column, there are local circuits that, too, are layer-specific. Thus, laminae ensure functional specificity of both afferent-target connections and local circuit connections.A prominent principle, which guides the formation of connections between specific cell types and is a characteristic feature of CNS architecture, is the concentration of synapses in small areas. These synapse clusters can take the form of planar layers or spherical glomeruli. Although glomeruli are a specialization that appears most prominent in the olfactory system, layers, or laminae, are an almost ubiquitous feature of central nervous system architecture. Indeed, even crude histological stains reveal that axons and dendrites often accumulate in neuropil (cell-body-free areas). Cell-type-specific or single-cell labeling has shown that, within individual neuropil layers, neurites and synapses are not distributed randomly. Rather, synaptic connections arising between neurons with the same or similar functional properties are localized to particular sublaminae that distinguish synapses with different properties (Fig. 1). The structural underpinnings of this functional principle are provided by mechanisms that ensure the lamina-specific branching of the corresponding neurites. How this enormous precision is achieved is the subject of intense investigations in the Drosophila, zebrafish, chick, and mouse visual systems. We will begin by describing three anatomical regions in these model organisms. Then, we will discuss three broad principles of layer-specific targeting in the visual system, namely cell–cell recognition, guidance by matrix cues, and activity-dependent sorting of axon terminals.  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