首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Myosin II heavy chain (MHC)-specific protein kinase C (MHC-PKC) isolated from the ameba, Dictyostelium discoideum, regulates myosin II assembly and localization in response to the chemoattractant cAMP (Abu- Elneel et al. 1996. J. Biol. Chem. 271:977- 984). Recent studies have indicated that cAMP-induced cGMP accumulation plays a role in the regulation of myosin II phosphorylation and localization (Liu, G., and P. Newell. 1991. J. Cell. Sci. 98: 483-490). This report describes the roles of cAMP and cGMP in the regulation of MHC-PKC membrane association, phosphorylation, and activity (hereafter termed MHC-PKC activities). cAMP stimulation of Dictyostelium cells resulted in translocation of MHC-PKC from the cytosol to the membrane fraction, as well as increasing in MHC-PKC phosphorylation and in its kinase activity. We present evidence that MHC is phosphorylated by MHC-PKC in the cell cortex which leads to myosin II dissociation from the cytoskeleton. Use of Dictyostelium mutants that exhibit aberrant cAMP- induced increases in cGMP accumulation revealed that MHC-PKC activities are regulated by cGMP. Dictyostelium streamer F mutant (stmF), which produces a prolonged peak of cGMP accumulation upon cAMP stimulation, exhibits prolonged increases in MHC-PKC activities. In contrast, Dictyostelium KI-10 mutant that lacks the normal cAMP-induced cGMP response, or KI-4 mutant that shows nearly normal cAMP-induced cGMP response but has aberrant cGMP binding activity, show no changes in MHC- PKC activities. We provide evidence that cGMP may affect MHC-PKC activities via the activation of cGMP-dependent protein kinase which, in turn, phosphorylates MHC-PKC. The results presented here indicate that cAMP-induced cGMP accumulation regulates myosin II phosphorylation and localization via the regulation of MHC-PKC.  相似文献   

2.
Chemotaxis-competent cells respond to a variety of ligands by activating second messenger pathways leading to changes in the actin/myosin cytoskeleton and directed cell movement. We demonstrate that Dictyostelium Akt/PKB, a homologue of mammalian Akt/PKB, is very rapidly and transiently activated by the chemoattractant cAMP. This activation takes place through G protein-coupled chemoattractant receptors via a pathway that requires homologues of mammalian p110 phosphoinositide-3 kinase. pkbA null cells exhibit aggregation-stage defects that include aberrant chemotaxis, a failure to polarize properly in a chemoattractant gradient and aggregation at low densities. Mechanistically, we demonstrate that the PH domain of Akt/PKB fused to GFP transiently translocates to the plasma membrane in response to cAMP with kinetics similar to those of Akt/PKB kinase activation and is localized to the leading edge of chemotaxing cells in vivo. Our results indicate Akt/PKB is part of the regulatory network required for sensing and responding to the chemoattractant gradient that mediates chemotaxis and aggregation.  相似文献   

3.
cAMP receptors mediate signal transduction pathways during development in Dictyostelium. A cAMP receptor (cAR1) has been cloned and sequenced (Klein, P., Sun, T. J., Saxe, C. L., Kimmel, A. R., Johnson, R. L., and Devreotes, P. N. (1988) Science 241, 1467-1472) and recently several other cAR genes have been identified (Saxe, C. L., Johnson, R., Devreotes, P. N., and Kimmel, A. R. (1991a) Dev. Genet. 12, 6-13; Saxe, C. L., Johnson, R. L., Devreotes, P. N., and Kimmel, A. R. (1991b) Genes Dev. 5, 1-8). We have expressed three receptor subtypes, cAR1, cAR2, and cAR3, in growing cells and have investigated their affinity and pharmacological specificity in a series of [3H]cAMP binding studies. In phosphate buffer, there were two affinity states of about 30 and 300 nM for cAR1 and 20 and 500 nM for cAR3 but no detectable affinity for cAR2. In the presence of 3 M ammonium sulfate, there was one affinity state of 4 nM for cAR1 and 11 nM for cAR2 and two affinity states of approximately 4 and 200 nM for cAR3. The relative affinities of 14 cyclic nucleotide derivatives were tested for each cAR in ammonium sulfate. These studies suggest a model (Van Haastert, P. J. M., and Kien, E. (1983) J. Biol. Chem. 258, 9636-9642) in which cAMP binds to all three receptor subtypes by maintaining hydrogen bond interactions at the N6 and O3' positions. Interactions at the exocyclic oxygens of cAMP varied between the receptors; cAR2 and cAR3 lacked a stereoselective interaction at the axial oxygen which was present in cAR1. The cleft, which binds the adenine ring of cAMP, was hydrophobic in cAR1 and cAR3 but relatively polar in cAR2. The analog specificity of cAR1 and cAR3 in phosphate buffer was similar to that measured in ammonium sulfate though the derivatives' relative affinity to cAMP was reduced. We conclude that these cAMP receptor subtypes can be distinguished by distinct pharmacological properties which will allow selective activation of each cAR during development.  相似文献   

4.
A key problem of eukaryotic cell motility is the signaling mechanism of chemoattractant gradient sensing. Recent experiments have revealed the molecular correlate of gradient sensing: Frontness molecules, such as PI3P and Rac, localize at the front end of the cell, and backness molecules, such as Rho and myosin II, accumulate at the back of the cell. Importantly, this frontness-backness polarization occurs spontaneously even if the cells are exposed to uniform chemoattractant profiles. The spontaneous polarization suggests that the gradient sensing machinery undergoes a Turing bifurcation. This has led to several classical activator-inhibitor and activator-substrate models which identify the frontness molecules with the activator. Conspicuously absent from these models is any accounting of the backness molecules. This stands in sharp contrast to experiments which show that the backness pathways inhibit the frontness pathways. Here, we formulate a model based on the mutually inhibitory interaction between the frontness and backness pathways. The model builds upon the mutual inhibition model proposed by Bourne and coworkers [Xu et al., 2003. Divergent signals and cytoskeletal assemblies regulate self-organizing polarity in neutrophils. Cell 114, 201-214.]. We show that mutual inhibition alone, without the help of any positive feedback (autocatalysis), can trigger spontaneous polarization of the frontness and backness pathways. The spatial distribution of the frontness and backness molecules in response to inhibition and activation of the frontness and backness pathways are consistent with those observed in experiments. Furthermore, depending on the parameter values, the model yields spatial distributions corresponding to chemoattraction (frontness pathways in-phase with the external gradient) and chemorepulsion (frontness pathways out-of-phase with the external gradient). Analysis of the model suggests a mechanism for the chemorepulsion-to-chemoattraction transition observed in neurons.  相似文献   

5.
We have previously reported the detection of cAMP waves within monolayers of aggregating Dictyostelium discoideum cells (K. J. Tomchik and P.N. Devreotes, 1981, Science 212, 443-446). The computer-assisted analysis presented here of the fluorographic images of the cAMP waves reveals (1) all the waves have a consistent width and height; (2) cAMP concentrations within centers of concentric aggregation territories oscillate periodically while at spiral centers the concentration builds up to a plateau value within 2 mm; (3) cells within the region of intersection of two oppositely directed cAMP waves are stimulated to produce more cAMP than those responding to a single wave; (4) cells start to move when the cAMP level begins to increase and cease movement when the peak cAMP concentration reaches the cell.  相似文献   

6.
《The Journal of cell biology》1995,129(6):1659-1665
We have previously reported that activation of adenylyl cyclase by chemoattractant receptors in Dictyostelium requires, in addition to a heterotrimeric G-protein, a cytosolic protein, designated CRAC (Lilly, P., and P. N. Devreotes. 1994. J. Biol. Chem. 269:14123-14129; Insall, R. H., A. Kuspa, P. J. Lilly, G. Schaulsky, L. R. Levin, W. F. Loomis, and P. N. Devreotes. 1994. J. Cell Biol. 126:1537-1545). In this report, we show that in intact cells, chemoattractants promote translocation of CRAC from the cytosolic to the membrane fraction. However, CRAC is not required at the time of receptor stimulation; it can be added to lysates of activated cells. Treatment of membranes with guanine nucleotides creates binding sites for CRAC. These binding sites can be generated in mutants lacking each of the components of the pathway except the beta-subunit, suggesting that free or "activated" beta gamma-subunits may be a part of the binding site. This hypothesis is consistent with previous observations that CRAC contains a pleckstrin homology domain and that the beta gamma-subunits likely mediate activation of adenylyl cyclase in this system. Thus, CRAC may serve as an adapter, linking the G-protein beta gamma-subunits to activation of the enzyme. GTP gamma S cannot generate CRAC-binding sites when the adenylyl cyclase pathway has been adapted by prior chemoattractant stimulation, suggesting that this is a point of downstream adaptation.  相似文献   

7.
Myosin II heavy chain (MHC) specific protein kinase C (MHC-PKC), isolated from Dictyostelium discoideum, regulates myosin II assembly and localization in response to the chemoattractant cyclic AMP. Immunoprecipitation of MHC-PKC revealed that it resides as a complex with several proteins. We show herein that one of these proteins is a homologue of the 14–3-3 protein (Dd14–3-3). This protein has recently been implicated in the regulation of intracellular signaling pathways via its interaction with several signaling proteins, such as PKC and Raf-1 kinase. We demonstrate that the mammalian 14–3-3 ζ isoform inhibits the MHC-PKC activity in vitro and that this inhibition is carried out by a direct interaction between the two proteins. Furthermore, we found that the cytosolic MHC-PKC, which is inactive, formed a complex with Dd14–3-3 in the cytosol in a cyclic AMP-dependent manner, whereas the membrane-bound active MHC-PKC was not found in a complex with Dd14–3-3. This suggests that Dd14–3-3 inhibits the MHC-PKC in vivo. We further show that MHC-PKC binds Dd14–3-3 as well as 14–3-3ζ through its C1 domain, and the interaction between these two proteins does not involve a peptide containing phosphoserine as was found for Raf-1 kinase. Our experiments thus show an in vivo function for a member of the 14–3-3 family and demonstrate that MHC-PKC interacts directly with Dd14–3-3 and 14–3-3ζ through its C1 domain both in vitro and in vivo, resulting in the inhibition of the kinase.  相似文献   

8.
In the simple amoeba Dictyostelium discoideum, myosin II filament assembly is regulated primarily by the action of a set of myosin heavy chain (MHC) kinases and by MHC phosphatase activity. Chemoattractant signals acting via G-protein coupled receptors lead to rapid recruitment of myosin II to the cell cortex, but the structural determinants on myosin necessary for translocation and the second messengers upstream of MHC kinases and phosphatases are not well understood. We report here the use of GFP-myosin II fusions to characterize the domains necessary for myosin II filament assembly and cytoskeletal recruitment during responses to global stimulation with the developmental chemoattractant cAMP. Analysis performed with GFP-myosin fusions, and with latrunculin A-treated cells, demonstrated that F-actin binding via the myosin motor domain together with concomitant filament assembly mediates the rapid cortical translocation observed in response to chemoattractant stimulation. A "headless" GFP-myosin construct lacking the motor domain was unable to translocate to the cell cortex in response to chemoattractant stimulation, suggesting that myosin motor-based motility may drive translocation. This lack of localization contrasts with previous work demonstrating accumulation of the same construct in the cleavage furrow of dividing cells, suggesting that recruitment signals and interactions during cytokinesis differ from those during chemoattractant responses. Evaluating upstream signaling, we find that iplA null mutants, devoid of regulated calcium fluxes during chemoattractant stimulation, display full normal chemoattractant-stimulated myosin assembly and translocation. These results indicate that calcium transients are not necessary for chemoattractant-regulated myosin II filament assembly and translocation.  相似文献   

9.
Directed cell migration occurs in response to extracellular cues. Following stimulation of a cell with chemoattractant, a significant rearrangement of the actin cytoskeleton is mediated by intracellular signaling pathways and results in polarization of the cell and movement via pseudopod extension. Amoeboid myosin Is play a critical role in regulating pseudopod formation in Dictyostelium, and their activity is activated by heavy chain phosphorylation. The effect of chemotactic stimulation on the in vivo phosphorylation level of a Dictyostelium myosin I, myoB, was tested. The myoB heavy chain is phosphorylated in vivo on serine 322 (the myosin TEDS rule phosphorylation site) in chemotactically competent cells. The level of myoB phosphorylation increases following stimulation of starving cells with the chemoattractant cAMP. A 3-fold peak increase in the level of phosphorylation is observed at 60 s following stimulation, a time at which the Dictyostelium cell actively extends pseudopodia. These findings suggest that chemotactic stimulation results in increased myoB activity via heavy chain phosphorylation and contributes to the global extension of pseudopodia that occurs prior to polarization and directed motility.  相似文献   

10.
Activation of G-protein-coupled chemoattractant receptors triggers dissociation of Galpha and Gbetagamma subunits. These subunits induce intracellular responses that can be highly polarized when a cell experiences a gradient of chemoattractant. Exactly how a cell achieves this amplified signal polarization is still not well understood. Here, we quantitatively measure temporal and spatial changes of receptor occupancy, G-protein activation by FRET imaging, and PIP3 levels by monitoring the dynamics of PH(Crac)-GFP translocation in single living cells in response to different chemoattractant fields. Our results provided the first direct evidence that G-proteins are activated to different extents on the cell surface in response to asymmetrical stimulations. A stronger, uniformly applied stimulation triggers not only a stronger G-protein activation but also a faster adaptation of downstream responses. When naive cells (which have not experienced chemoattractant) were abruptly exposed to stable cAMP gradients, G-proteins were persistently activated throughout the entire cell surface, whereas the response of PH(Crac)-GFP translocation surprisingly consisted of two phases, an initial transient and asymmetrical translocation around the cell membrane, followed by a second phase producing a highly polarized distribution of PH(Crac)-GFP. We propose a revised model of gradient sensing, suggesting an important role for locally controlled components that inhibit PI3Kinase activity.  相似文献   

11.
We have investigated the role of Rap1 in controlling chemotaxis and cell adhesion in Dictyostelium discoideum. Rap1 is activated rapidly in response to chemoattractant stimulation, and activated Rap1 is preferentially found at the leading edge of chemotaxing cells. Cells expressing constitutively active Rap1 are highly adhesive and exhibit strong chemotaxis defects, which are partially caused by an inability to spatially and temporally regulate myosin assembly and disassembly. We demonstrate that the kinase Phg2, a putative Rap1 effector, colocalizes with Rap1-guanosine triphosphate at the leading edge and is required in an in vitro assay for myosin II phosphorylation, which disassembles myosin II and facilitates filamentous actin-mediated leading edge protrusion. We suggest that Rap1/Phg2 plays a role in controlling leading edge myosin II disassembly while passively allowing myosin II assembly along the lateral sides and posterior of the cell.  相似文献   

12.
Chemotaxis, the directed migration of leukocytes towards a chemoattractant gradient, is a key phenomenon in the immune response. During lymphocyte-endothelial and – extracellular matrix interactions, chemokines induce the polarization of T lymphocytes. with generation of specialized cell compartments. The chemokine receptors involved in detection of the chemoattractant gradients concentrate at the leading edge (advancing front or anterior pole) of the cell. The adhesion molecules ICAM- 1, -3, CD44 and CD43 redistribute to the uropod, an appendage at the posterior pole of migrating T lymphocyte that protrudes from the contact area with endothelial or extracellular matrix substrates. Whereas chemokine receptors sense the direction of migration, the uropod is involved in the recruitment of bystander leukocytes through LFA-1/ICAM-dependent cell cell interactions. While β-actin concentrates preferentially at the cell's leading edge, the motor protein myosin II and a microtubule organizing center (MTOC) are packed in the uropod. The actin-binding protein moesin, which belongs to the ERM family of ezrin, radixin and moesin, redistributes to the distal portion of uropods and physically interacts with ICAM-3, CD44 and CD43, thus acting as a physical link between the membrane molecules and the actin cytoskeleton. Moreover, the moesin-ICAM-3 association correlates with the degree of cell polarity. The redistribution of the chemokine receptors and adhesion molecules to opposite poles of the cell in response to a chemoattractant gradient may guide cell migration and cell-cell interactions during lymphoid cell trafficking in immune and inflammatory responses.  相似文献   

13.
14.
Previous results, based on inhibition of fertilization by an anti-alpha6 integrin mAb (GoH3), suggest that the alpha6beta1 integrin on mouse eggs functions as the receptor for sperm (Almeida, E.A., A.P. Huovila, A.E. Sutherland, L.E. Stephens, P.G. Calarco, L. M. Shaw, A.M. Mercurio, A. Sonnenberg, P. Primakoff, D.G. Myles, and J.M. White. 1995. Cell. 81:1095-1104). Because the egg surface tetraspanin CD9 is essential for gamete fusion (Kaji, K., S. Oda, T. Shikano, T. Ohnuki, Y. Uematsu, J. Sakagami, N. Tada, S. Miyazaki, and A. Kudo. 2000. Nat. Genet. 24:279-282; Le Naour, F., E. Rubinstein, C. Jasmin, M. Prenant, and C. Boucheix. 2000. Science. 287:319-321; Miyado, K., G. Yamada, S. Yamada, H. Hasuwa, Y. Nakamura, F. Ryu, K. Suzuki, K. Kosai, K. Inoue, A. Ogura, M. Okabe, and E. Mekada. 2000. Science. 287:321-324) and CD9 is known to associate with integrins, recent models of gamete fusion have posited that egg CD9 acts in association with alpha6beta1 in fusion (Chen, M.S., K.S. Tung, S.A. Coonrod, Y. Takahashi, D. Bigler, A. Chang, Y. Yamashita, P.W. Kincade, J.C. Herr, and J.M. White. 1999. Proc. Natl. Acad. Sci. USA. 96:11830-11835; Kaji, K., S. Oda, T. Shikano, T. Ohnuki, Y. Uematsu, J. Sakagami, N. Tada, S. Miyazaki, and A. Kudo. 2000. Nat. Genet. 24:279-282; Le Naour, F., E. Rubinstein, C. Jasmin, M. Prenant, and C. Boucheix. 2000. Science. 287:319-321; Miyado, K., G. Yamada, S. Yamada, H. Hasuwa, Y. Nakamura, F. Ryu, K. Su- zuki, K. Kosai, K. Inoue, A. Ogura, M. Okabe, and E. Mekada. 2000. Science. 287:321-324). Using eggs from cultured ovaries of mice lacking the alpha6 integrin subunit, we found that the fertilization rate, fertilization index, and sperm binding were not impaired compared with wild-type or heterozygous controls. Furthermore, a reexamination of antibody inhibition, using an assay that better simulates in vivo fertilization conditions, revealed no inhibition of fusion by the GoH3 mAb. We also found that an anti-CD9 mAb completely blocks sperm fusion with either wild-type eggs or eggs lacking alpha6beta1. Based on these results, we conclude that the alpha6beta1 integrin is not essential for sperm-egg fusion, and we suggest a new model in which CD9 acts by itself, or interacts with egg protein(s) other than alpha6beta1, to function in sperm-egg fusion.  相似文献   

15.
Increased intracellular H(+) efflux is speculated to be an evolutionarily conserved mechanism necessary for rapid assembly of cytoskeletal filaments and for morphological polarity during cell motility. In Dictyostelium discoideum, increased intracellular pH through undefined transport mechanisms plays a key role in directed cell movement. We report that a developmentally regulated Na-H exchanger in Dictyostelium discoideum (DdNHE1) localizes to the leading edge of polarized cells and is necessary for intracellular pH homeostasis and for efficient chemotaxis. Starved DdNHE1-null cells (Ddnhe1(-)) differentiate, and in response to the chemoattractant cAMP they retain directional sensing; however, they cannot attain a polarized morphology, but instead extend mislocalized pseudopodia around the cell and exhibit decreased velocity. Consistent with impaired polarity, in response to chemoattractant, Ddnhe1(-) cells lack a leading edge localization of F-actin and have significantly attenuated de novo F-actin polymerization but increased abundance of membrane-associated phosphatidylinositol 3,4,5-trisphosphate (PI((3,4,5))P(3)). These findings indicate that during chemotaxis DdNHE1 is necessary for establishing the kinetics of actin polymerization and PI((3,4,5))P(3) production and for attaining a polarized phenotype.  相似文献   

16.
Myosin II is not essential for cytokinesis in cells of Dictyostelium discoideum that are anchored on a substrate (Neujahr, R., C. Heizer, and G. Gerisch. 1997. J. Cell Sci. 110:123–137), in contrast to its importance for cell division in suspension (DeLozanne, A., and J.A. Spudich. 1987. Science. 236:1086–1091; Knecht, D.A., and W.F. Loomis. 1987. Science. 236: 1081–1085.). These differences have prompted us to investigate the three-dimensional distribution of myosin II in cells dividing under one of three conditions: (a) in shaken suspension, (b) in a fluid layer on a solid substrate surface, and (c) under mechanical stress applied by compressing the cells. Under the first and second conditions outlined above, myosin II does not form patterns that suggest a contractile ring is established in the furrow. Most of the myosin II is concentrated in the regions that flank the furrow on both sides towards the poles of the dividing cell. It is only when cells are compressed that myosin II extensively accumulates in the cleavage furrow, as has been previously described (Fukui, Y., T.J. Lynch, H. Brzeska, and E.D. Korn. 1989. Nature. 341:328–331), i.e., this massive accumulation is a response to the mechanical stress. Evidence is provided that the stress-associated translocation of myosin II to the cell cortex is a result of the dephosphorylation of its heavy chains. F-actin is localized in the dividing cells in a distinctly different pattern from that of myosin II. The F-actin is shown to accumulate primarily in protrusions at the two poles that ultimately form the leading edges of the daughter cells. This distribution changes dynamically as visualized in living cells with a green fluorescent protein–actin fusion.  相似文献   

17.
A wound induces cell polarization, in which myosin II is localized at the rear end of individual cells in a migrating epithelial sheet of the Drosophila larval epidermis. Here, we use myosin localization to demonstrate that Rac1, Cdc42, and Rho1 are each required for cell polarization and directional sensing of the wound. The three GTPases are also required for actin cable formation at the wound leading edge. Rac1, Cdc42, and Rho1 act upstream of c-Jun N-terminal kinase (JNK) to organize actin assembly. These results highlight the similarities between the molecular mechanism of Drosophila wound healing and those of Drosophila embryonic dorsal closure and the chemotactic response of Dictyostelium and leukocytes.  相似文献   

18.
Role of phosphatidylinositol 3-kinases in chemotaxis in Dictyostelium   总被引:1,自引:0,他引:1  
Experiments in several cell types revealed that local accumulation of phosphatidylinositol 3,4,5-triphosphate mediates the ability of cells to migrate during gradient sensing. We took a systematic approach to characterize the functions of the six putative Class I phosphatidylinositol 3-kinases (PI3K1-6) in Dictyostelium by creating a series of gene knockouts. These studies revealed that PI3K1-PI3K3 are the major PI3Ks for chemoattractant-mediated phosphatidylinositol 3,4,5-triphosphate production. We studied chemotaxis of the pi3k1/2/3 triple knock-out strain (pi3k1/2/3 null cells) to cAMP under two distinct experimental conditions, an exponential gradient emitted from a micropipette and a shallow, linear gradient in a Dunn chamber, using four cAMP concentrations ranging over a factor of 10,000. Under all conditions tested pi3k1/2/3 null cells moved slower and had less polarity than wild-type cells. pi3k1/2/3 null cells moved toward a chemoattractant emitted by a micropipette, although persistence was lower than that of wild-type or pi3k1/2 null cells. In shallow linear gradients, pi3k1/2 null cells had greater directionality defects, especially at lower chemoattractant concentrations. Our studies suggest that although PI3K is not essential for directional movement under some chemoattractant conditions, it is a key component of the directional sensing pathway and plays a critical role in linear chemoattractant gradients, especially at low chemoattractant concentrations. The relative importance of PI3K in chemotaxis is also dependent on the developmental stage of the cells. Our data suggest that the output of other signaling pathways suffices to mediate directional sensing when cells perceive a strong signal, but PI3K signaling is crucial for detecting weaker signals.  相似文献   

19.
We have identified a Dictyostelium discoideum gene encoding a serine/threonine kinase, PAKa, a putative member of the Ste20/PAK family of p21-activated kinases, with a kinase domain and a long NH(2)-terminal regulatory domain containing an acidic segment, a polyproline domain, and a CRIB domain. PAKa colocalizes with myosin II to the cleavage furrow of dividing cells and the posterior of polarized, chemotaxing cells via its NH(2)-terminal domain. paka null cells are defective in completing cytokinesis in suspension. PAKa is also required for maintaining the direction of cell movement, suppressing lateral pseudopod extension, and proper retraction of the posterior of chemotaxing cells. paka null cells are defective in myosin II assembly, as the myosin II cap in the posterior of chemotaxing cells and myosin II assembly into cytoskeleton upon cAMP stimulation are absent in these cells, while constitutively active PAKa leads to an upregulation of myosin II assembly. PAKa kinase activity against histone 2B is transiently stimulated and PAKa incorporates into the cytoskeleton with kinetics similar to those of myosin II assembly in response to chemoattractant signaling. However, PAKa does not phosphorylate myosin II. We suggest that PAKa is a major regulator of myosin II assembly, but does so by negatively regulating myosin II heavy chain kinase.  相似文献   

20.
Dictyostelium discoideum cells have been generated that lack myosin heavy chain (MHC) due to antisense RNA inactivation of the endogenous mRNA or to insertional mutagenesis of the myosin gene. These cells retain chemotactic movement in gradients of the chemoattractant cAMP. Furthermore, cAMP does induce many biochemical and physiological responses in aggregative cells, including binding of cAMP to surface receptors, modification, and down-regulation of the receptor; activation of adenylate and guanylate cyclase, secretion of cAMP; and the association of actin to the Triton-insoluble cytoskeleton. Cells lacking MHC were found to have a requirement for bivalent cations in the medium for optimal chemotaxis and cell aggregation.  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