共查询到20条相似文献,搜索用时 46 毫秒
1.
Ellen J. Tisdale Fouad Azizi Cristina R. Artalejo 《The Journal of biological chemistry》2009,284(9):5876-5884
Rab2 requires glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and atypical
protein kinase Cι (aPKCι) for retrograde vesicle formation from
vesicular tubular clusters that sort secretory cargo from recycling proteins
returned to the endoplasmic reticulum. However, the precise role of GAPDH and
aPKCι in the early secretory pathway is unclear. GAPDH was the first
glycolytic enzyme reported to co-purify with microtubules (MTs). Similarly,
aPKC associates directly with MTs. To learn whether Rab2 also binds directly
to MTs, a MT binding assay was performed. Purified Rab2 was found in a
MT-enriched pellet only when both GAPDH and aPKCι were present, and
Rab2-MT binding could be prevented by a recombinant fragment made to the Rab2
amino terminus (residues 2-70), which directly interacts with GAPDH and
aPKCι. Because GAPDH binds to the carboxyl terminus of α-tubulin,
we characterized the distribution of tyrosinated/detyrosinated α-tubulin
that is recruited by Rab2 in a quantitative membrane binding assay.
Rab2-treated membranes contained predominantly tyrosinated α-tubulin;
however, aPKCι was the limiting and essential factor.
Tyrosination/detyrosination influences MT motor protein binding; therefore, we
determined whether Rab2 stimulated kinesin or dynein membrane binding.
Although kinesin was not detected on membranes incubated with Rab2, dynein was
recruited in a dose-dependent manner, and binding was aPKCι-dependent.
These combined results suggest a mechanism by which Rab2 controls MT and motor
recruitment to vesicular tubular clusters.The small GTPase Rab2 is essential for membrane trafficking in the early
secretory pathway and associates with vesicular tubular
clusters
(VTCs)2 located
between the endoplasmic reticulum (ER) and the cis-Golgi compartment
(1,
2). VTCs are pleomorphic
structures that sort anterograde-directed cargo from recycling proteins and
trafficking machinery retrieved to the ER
(3-6).
Rab2 bound to a VTC microdomain stimulates recruitment of soluble factors that
results in the release of vesicles containing the recycling protein p53/p58
(7). In that regard, we have
previously reported that glyceraldehyde-3-phosphate dehydrogenase (GAPDH) and
atypical PKC ι (aPKCι) are Rab2 effectors that interact directly
with the Rab2 amino terminus and with each other
(8,
9). Their interaction requires
Src-dependent tyrosine phosphorylation of GAPDH and aPKCι
(10). Moreover, GAPDH is a
substrate for aPKCι (11).
GAPDH catalytic activity is not required for ER to Golgi transport indicating
that GAPDH provides a specific function essential for membrane trafficking
from VTCs independent of glycolytic function
(9). Indeed, phospho-GAPDH
influences MT dynamics in the early secretory pathway
(11).GAPDH was the first glycolytic enzyme reported to co-purify with
microtubules (MTs) (12) and
subsequently was shown to interact with the carboxyl terminus of
α-tubulin (13). The
binding of GAPDH to MTs promotes formation of cross-linked parallel MT arrays
or bundles (14,
15). GAPDH has also been
reported to possess membrane fusogenic activity, which is inhibited by tubulin
(16). Similarly, aPKC
associates directly with tubulin and promotes MT stability and MT remodeling
at specific intracellular sites
(17-21).
It may not be coincidental that these two Rab2 effectors influence MT dynamics
because recent studies indicate that the cytoskeleton plays a central role in
the organization and operation of the secretory pathway
(22).MTs are dynamic structures that grow or shrink by the addition or loss of
α- and β-tubulin heterodimers from the ends of protofilaments
(23). Their assembly and
stability is regulated by a variety of proteins traditionally referred to as
microtubule-associated proteins (MAPs). In addition to the multiple
α/β isoforms that are present in eukaryotes, MTs undergo an
assortment of post-translational modifications, including acetylation,
glycylation, glutamylation, phosphorylation, palmitoylation, and
detyrosination, which further contribute to their biochemical heterogeneity
(24,
25). It has been proposed that
these tubulin modifications regulate intracellular events by facilitating
interaction with MAPs and with other specific effector proteins
(24). For example, the
reversible addition of tyrosine to the carboxyl terminus of α-tubulin
regulates MT interaction with plus-end tracking proteins (+TIPs) containing
the cytoskeleton-associated protein glycine-rich (CAP-Gly) motif and with
dynein-dynactin
(27-29).
Additionally, MT motility and cargo transport rely on the cooperation of the
motor proteins kinesin and dynein
(30). Kinesin is a plus-end
directed MT motor, whereas cytoplasmic dynein is a minus-end MT-based motor,
and therefore the motors transport vesicular cargo toward the opposite end of
a MT track (31).Although MT assembly does not appear to be directly regulated by small
GTPases, Rab proteins provide a molecular link for vesicle movement along MTs
to the appropriate target (22,
32-34).
In this study, the potential interaction of Rab2 with MTs and motor proteins
was characterized. We found that Rab2 does not bind directly to preassembled
MTs but does associate when both GAPDH and aPKCι are present and bound to
MTs. Moreover, the MTs predominantly contained tyrosinated α-tubulin
(Tyr-tubulin) suggesting that a dynamic pool of MTs that differentially binds
MAPs/effector proteins/motors associates with VTCs in response to Rab2. To
that end, we determined that Rab2-promoted dynein/dynactin binding to
membranes and that the recruitment required aPKCι. 相似文献
2.
Identification of the Gene Encoding the Tryptophan Synthase
β-Subunit from Chlamydomonas reinhardtii 下载免费PDF全文
We report the isolation of a Chlamydomonas reinhardtii cDNA that encodes the β-subunit of tryptophan synthase (TSB). This cDNA was cloned by functional complementation of a trp-operon-deleted strain of Escherichia coli. Hybridization analysis indicated that the gene exists in a single copy. The predicted amino acid sequence showed the greatest identity to TSB polypeptides from other photosynthetic organisms. With the goal of identifying mutations in the gene encoding this enzyme, we isolated 11 recessive and 1 dominant single-gene mutation that conferred resistance to 5-fluoroindole. These mutations fell into three complementation groups, MAA2, MAA7, and TAR1. In vitro assays showed that mutations at each of these loci affected TSB activity. Restriction fragment-length polymorphism analysis suggested that MAA7 encodes TSB. MAA2 and TAR1 may act to regulate the activity of MAA7 or its protein product. 相似文献
3.
4.
5.
The translocation and accumulation of glyceraldehyde-3-phosphate dehydrogenase (GAPDH) in the nucleus has closely been associated with cell death induction. However, the mechanism of this process has not been completely understood. The E3 ubiquitin ligase siah-1 (seven in absentia homolog 1) has recently been identified as a potential shuttle protein to transport GAPDH from the cytosol to the nucleus. Previously, we have demonstrated that elevated glucose levels induce GAPDH nuclear accumulation in retinal Müller cells. Therefore, this study investigated the role of siah-1 in high glucose-induced GAPDH nuclear translocation and subsequent cell death in retinal Müller cells. High glucose significantly increased siah-1 expression within 12 h. Under hyperglycemic conditions, siah-1 formed a complex with GAPDH and was predominantly localized in the nucleus of Müller cells. siah-1 knockdown using 50 nm siah-1 small interfering RNA significantly decreased high glucose-induced GAPDH nuclear accumulation at 24 h by 43.8 ± 4.0%. Further, knockdown of siah-1 prevented high glucose-induced cell death of Müller cells potentially by inhibiting p53 phosphorylation consistent with previous observations, indicating that nuclear GAPDH induces cell death via p53 activation. Therefore, inhibition of GAPDH nuclear translocation and accumulation by targeting siah-1 promotes Müller cell survival under hyperglycemic conditions. 相似文献
6.
IDDARAbdelghani CAMPOSLuisA SANCHOJavier SERRANOAurelio SOUKRIAbdelaziz 《Acta biochimica et biophysica Sinica》2003,35(10):891-896
D Glyceraldehyde 3 phosphatedehydrogenase(GAPDH ,EC 1.2 .1.12 )isakeyenzymeoftheglycolyticpathwaythatispresentinthecytosolofallorganismssofarstudied[1] .TheglycolyticGAPDHhasbeenremarkablyconservedduringevolution ,havingahomotetramericstructurewithsubunitsof 35 - 37kD[1] .GAPDHhasbeenisolatedfromavarietyofspecies[2 ] ,includingmesophilic ,moderatelythermophilicandhyperthermophilicmicroorganisms[3 ] .Theseenzymes ,whichdifferinthermalstability ,havebeenshowntobehighlysimilarinaminoacidse… 相似文献
7.
Mikael K. Schnizler Katrin Schnizler Xiang-ming Zha Duane D. Hall John A. Wemmie Johannes W. Hell Michael J. Welsh 《The Journal of biological chemistry》2009,284(5):2697-2705
The acid-sensing ion channel 1a (ASIC1a) is widely expressed in central and
peripheral neurons where it generates transient cation currents when
extracellular pH falls. ASIC1a confers pH-dependent modulation on postsynaptic
dendritic spines and has critical effects in neurological diseases associated
with a reduced pH. However, knowledge of the proteins that interact with
ASIC1a and influence its function is limited. Here, we show that
α-actinin, which links membrane proteins to the actin cytoskeleton,
associates with ASIC1a in brain and in cultured cells. The interaction
depended on an α-actinin-binding site in the ASIC1a C terminus that was
specific for ASIC1a versus other ASICs and for α-actinin-1 and
-4. Co-expressing α-actinin-4 altered ASIC1a current density, pH
sensitivity, desensitization rate, and recovery from desensitization.
Moreover, reducing α-actinin expression altered acid-activated currents
in hippocampal neurons. These findings suggest that α-actinins may link
ASIC1a to a macromolecular complex in the postsynaptic membrane where it
regulates ASIC1a activity.Acid-sensing ion channels
(ASICs)2 are
H+-gated members of the DEG/ENaC family
(1–3).
Members of this family contain cytosolic N and C termini, two transmembrane
domains, and a large cysteine-rich extracellular domain. ASIC subunits combine
as homo- or heterotrimers to form cation channels that are widely expressed in
the central and peripheral nervous systems
(1–4).
In mammals, four genes encode ASICs, and two subunits, ASIC1 and ASIC2, have
two splice forms, a and b. Central nervous system neurons express ASIC1a,
ASIC2a, and ASIC2b
(5–7).
Homomeric ASIC1a channels are activated when extracellular pH drops below 7.2,
and half-maximal activation occurs at pH 6.5–6.8
(8–10).
These channels desensitize in the continued presence of a low extracellular
pH, and they can conduct Ca2+
(9,
11–13).
ASIC1a is required for acid-evoked currents in central nervous system neurons;
disrupting the gene encoding ASIC1a eliminates H+-gated currents
unless extracellular pH is reduced below pH 5.0
(5,
7).Previous studies found ASIC1a enriched in synaptosomal membrane fractions
and present in dendritic spines, the site of excitatory synapses
(5,
14,
15). Consistent with this
localization, ASIC1a null mice manifested deficits in hippocampal
long term potentiation, learning, and memory, which suggested that ASIC1a is
required for normal synaptic plasticity
(5,
16). ASICs might be activated
during neurotransmission when synaptic vesicles empty their acidic contents
into the synaptic cleft or when neuronal activity lowers extracellular pH
(17–19).
Ion channels, including those at the synapse often interact with multiple
proteins in a macromolecular complex that incorporates regulators of their
function (20,
21). For ASIC1a, only a few
interacting proteins have been identified. Earlier work indicated that ASIC1a
interacts with another postsynaptic scaffolding protein, PICK1
(15,
22,
23). ASIC1a also has been
reported to interact with annexin II light chain p11 through its cytosolic N
terminus to increase cell surface expression
(24) and with
Ca2+/calmodulin-dependent protein kinase II to phosphorylate the
channel (25). However, whether
ASIC1a interacts with additional proteins and with the cytoskeleton remain
unknown. Moreover, it is not known whether such interactions alter ASIC1a
function.In analyzing the ASIC1a amino acid sequence, we identified cytosolic
residues that might bind α-actinins. α-Actinins cluster membrane
proteins and signaling molecules into macromolecular complexes and link
membrane proteins to the actincytoskeleton (for review, Ref.
26). Four genes encode
α-actinin-1, -2, -3, and -4 isoforms. α-Actinins contain an
N-terminal head domain that binds F-actin, a C-terminal region containing two
EF-hand motifs, and a central rod domain containing four spectrin-like motifs
(26–28).
The C-terminal portion of the rod segment appears to be crucial for binding to
membrane proteins. The α-actinins assemble into antiparallel homodimers
through interactions in their rod domain. α-Actinins-1, -2, and -4 are
enriched in dendritic spines, concentrating at the postsynaptic membrane
(29–35).
In the postsynaptic membrane of excitatory synapses, α-actinin connects
the NMDA receptor to the actin cytoskeleton, and this interaction is key for
Ca2+-dependent inhibition of NMDA receptors
(36–38).
α-Actinins can also regulate the membrane trafficking and function of
several cation channels, including L-type Ca2+ channels,
K+ channels, and TRP channels
(39–41).To better understand the function of ASIC1a channels in macromolecular
complexes, we asked if ASIC1a associates with α-actinins. We were
interested in the α-actinins because they and ASIC1a, both, are present
in dendritic spines, ASIC1a contains a potential α-actinin binding
sequence, and the related epithelial Na+ channel (ENaC) interacts
with the cytoskeleton (42,
43). Therefore, we
hypothesized that α-actinin interacts structurally and functionally with
ASIC1a. 相似文献
8.
Yamini S. Bynagari Bela Nagy Jr. Florin Tuluc Kamala Bhavaraju Soochong Kim K. Vinod Vijayan Satya P. Kunapuli 《The Journal of biological chemistry》2009,284(20):13413-13421
The novel class of protein kinase C (nPKC) isoform η is expressed in
platelets, but not much is known about its activation and function. In this
study, we investigated the mechanism of activation and functional implications
of nPKCη using pharmacological and gene knock-out approaches. nPKCη
was phosphorylated (at Thr-512) in a time- and concentration-dependent manner
by 2MeSADP. Pretreatment of platelets with MRS-2179, a P2Y1
receptor antagonist, or YM-254890, a Gq blocker, abolished
2MeSADP-induced phosphorylation of nPKCη. Similarly, ADP failed to
activate nPKCη in platelets isolated from P2Y1 and
Gq knock-out mice. However, pretreatment of platelets with
P2Y12 receptor antagonist, AR-C69331MX did not interfere with
ADP-induced nPKCη phosphorylation. In addition, when platelets were
activated with 2MeSADP under stirring conditions, although nPKCη was
phosphorylated within 30 s by ADP receptors, it was also dephosphorylated by
activated integrin αIIbβ3 mediated outside-in
signaling. Moreover, in the presence of SC-57101, a
αIIbβ3 receptor antagonist, nPKCη
dephosphorylation was inhibited. Furthermore, in murine platelets lacking
PP1cγ, a catalytic subunit of serine/threonine phosphatase,
αIIbβ3 failed to dephosphorylate nPKCη.
Thus, we conclude that ADP activates nPKCη via P2Y1 receptor
and is subsequently dephosphorylated by PP1γ phosphatase activated by
αIIbβ3 integrin. In addition, pretreatment of
platelets with η-RACK antagonistic peptides, a specific inhibitor of
nPKCη, inhibited ADP-induced thromboxane generation. However, these
peptides had no affect on ADP-induced aggregation when thromboxane generation
was blocked. In summary, nPKCη positively regulates agonist-induced
thromboxane generation with no effects on platelet aggregation.Platelets are the key cellular components in maintaining hemostasis
(1). Vascular injury exposes
subendothelial collagen that activates platelets to change shape, secrete
contents of granules, generate thromboxane, and finally aggregate via
activated αIIbβ3 integrin, to prevent further
bleeding (2,
3). ADP is a physiological
agonist of platelets secreted from dense granules and is involved in feedback
activation of platelets and hemostatic plug stabilization
(4). It activates two distinct
G-protein-coupled receptors (GPCRs) on platelets, P2Y1 and
P2Y12, which couple to Gq and Gi,
respectively
(5–8).
Gq activates phospholipase Cβ (PLCβ), which leads to
diacyl glycerol (DAG)2
generation and calcium mobilization
(9,
10). On the other hand,
Gi is involved in inhibition of cAMP levels and PI 3-kinase
activation (4,
6). Synergistic activation of
Gq and Gi proteins leads to the activation of the
fibrinogen receptor integrin αIIbβ3.
Fibrinogen bound to activated integrin αIIbβ3
further initiates feed back signaling (outside-in signaling) in platelets that
contributes to the formation of a stable platelet plug
(11).Protein kinase Cs (PKCs) are serine/threonine kinases known to regulate
various platelet functional responses such as dense granule secretion and
integrin αIIbβ3 activation
(12,
13). Based on their structure
and cofactor requirements, PKCs are divided in to three classes: classical
(cofactors: DAG, Ca2+), novel (cofactors: DAG) and atypical
(cofactors: PIP3) PKC isoforms
(14). All the members of the
novel class of PKC isoforms (nPKC), viz. nPKC isoforms δ, θ,
η, and ε, are expressed in platelets
(15), and they require DAG for
activation. Among all the nPKCs, PKCδ
(15,
16) and PKCθ
(17–19)
are fairly studied in platelets. Whereas nPKCδ is reported to regulate
protease-activated receptor (PAR)-mediated dense granule secretion
(15,
20), nPKCθ is activated
by outside-in signaling and contributes to platelet spreading on fibrinogen
(18). On the other hand, the
mechanism of activation and functional role of nPKCη is not addressed as
yet.PKCs are cytoplasmic enzymes. The enzyme activity of PKCs is modulated via
three mechanisms (14,
21): 1) cofactor binding: upon
cell stimulus, cytoplasmic PKCs mobilize to membrane, bind cofactors such as
DAG, Ca2+, or PIP3, release autoinhibition, and attain an active
conformation exposing catalytic domain of the enzyme. 2) phosphorylations:
3-phosphoinositide-dependent kinase 1 (PDK1) on the membrane phosphorylates
conserved threonine residues on activation loop of catalytic domain; this is
followed by autophosphorylations of serine/threonine residues on turn motif
and hydrophobic region. These series of phosphorylations maintain an active
conformation of the enzyme. 3) RACK binding: PKCs in active conformation bind
receptors for activated C kinases (RACKs) and are lead to various subcellular
locations to access the substrates
(22,
23). Although various leading
laboratories have elucidated the activation of PKCs, the mechanism of
down-regulation of PKCs is not completely understood.The premise of dynamic cell signaling, which involves protein
phosphorylations by kinases and dephosphorylations by phosphatases has gained
immense attention over recent years. PP1, PP2A, PP2B, PHLPP are a few of the
serine/threonine phosphatases reported to date. Among them PP1 and PP2
phosphatases are known to regulate various platelet functional responses
(24,
25). Furthermore, PP1c, is the
catalytic unit of PP1 known to constitutively associate with
αIIb and is activated upon integrin engagement with
fibrinogen and subsequent outside-in signaling
(26). Among various PP1
isoforms, recently PP1γ is shown to positively regulate platelet
functional responses (27).
Thus, in this study we investigated if the above-mentioned phosphatases are
involved in down-regulation of nPKCη. Furthermore, reports from other cell
systems suggest that nPKCη regulates ERK/JNK pathways
(28). In platelets ERK is
known to regulate agonist induced thromboxane generation
(29,
30). Thus, we also
investigated if nPKCη regulates ERK phosphorylation and thereby
agonist-induced platelet functional responses.In this study, we evaluated the activation of nPKCη downstream of ADP
receptors and its inactivation by an integrin-associated phosphatase
PP1γ. We also studied if nPKCη regulates functional responses in
platelets and found that this isoform regulates ADP-induced thromboxane
generation, but not fibrinogen receptor activation in platelets. 相似文献
9.
《BBA》1987,893(1):91-99
Analyses of chlorophyll-protein complexes and of lipids were performed with the wild type of Chlamydomonas reinhardtii and three non-photosynthetic mutants: Fl 39, which was a ‘classical’ high-fluorescent Photosystem II (PS II)-lacking mutant, and mf 1 and mf 2, which lacked also functional PS II but were low-fluorescent and showed an abnormally predominant energy transfer from the main light-harvesting antenna towards Photosystem I. An oligomeric form of the chlorophyll a + b-protein complex CP II was clearly isolated from the wild type and the mutant Fl 39 but it was not detected in the mutants mf 1 and mf 2. The three mutants showed total lipid contents close to or greater than that of the wild type. Their phosphatidyldiacylglycerol (PG) contents, on a chlorophyll basis, were higher (Fl 39) or 1.4- (mf 1) and 2.0- (mf 2) times lower than that of the wild type. The fatty acid compositions of the wild type and of the mutant Fl 39 were comparable, showing about equal amounts of a C18 series and a C16 series which included the Δ3-trans-hexadecenoic acid (C16:1-trans). This C16:1-trans was not detected in the mutants mf 1 and mf 2 which contained the other fatty acids. These results indicate correlations between lack of C16:1-trans-containing PG, lack of an oligomeric form of CP II and an impaired mechanism of the regulation of excitation energy transfer from the main chlorophyll a + b antenna. 相似文献
10.
Annamari Paino Tuuli Ahlstrand Jari Nuutila Indre Navickaite Maria Lahti Heidi Tuominen Hannamari V?limaa Urpo Lamminm?ki Marja T. P?ll?nen Riikka Ihalin 《PloS one》2013,8(7)
Aggregatibacter
actinomycetemcomitans
is a gram-negative opportunistic oral pathogen. It is frequently associated with subgingival biofilms of both chronic and aggressive periodontitis, and the diseased sites of the periodontium exhibit increased levels of the proinflammatory mediator interleukin (IL)-1β. Some bacterial species can alter their physiological properties as a result of sensing IL-1β. We have recently shown that this cytokine localizes to the cytoplasm of A. actinomycetemcomitans in co-cultures with organotypic gingival mucosa. However, current knowledge about the mechanism underlying bacterial IL-1β sensing is still limited. In this study, we characterized the interaction of A. actinomycetemcomitans total membrane protein with IL-1β through electrophoretic mobility shift assays. The interacting protein, which we have designated bacterial interleukin receptor I (BilRI), was identified through mass spectrometry and was found to be Pasteurellaceae specific. Based on the results obtained using protein function prediction tools, this protein localizes to the outer membrane and contains a typical lipoprotein signal sequence. All six tested biofilm cultures of clinical A. actinomycetemcomitans strains expressed the protein according to phage display-derived antibody detection. Moreover, proteinase K treatment of whole A. actinomycetemcomitans cells eliminated BilRI forms that were outer membrane specific, as determined through immunoblotting. The protein was overexpressed in Escherichia coli in both the outer membrane-associated form and a soluble cytoplasmic form. When assessed using flow cytometry, the BilRI-overexpressing E. coli cells were observed to bind 2.5 times more biotinylated-IL-1β than the control cells, as detected with avidin-FITC. Overexpression of BilRI did not cause binding of a biotinylated negative control protein. In a microplate assay, soluble BilRI bound to IL-1β, but this binding was not specific, as a control protein for IL-1β also interacted with BilRI. Our findings suggest that A. actinomycetemcomitans expresses an IL-1β-binding surface-exposed lipoprotein that may be part of the bacterial IL-1β-sensing system. 相似文献
11.
Yongmei Pu Susan H. Garfield Noemi Kedei Peter M. Blumberg 《The Journal of biological chemistry》2009,284(2):1302-1312
Classic and novel protein kinase C (PKC) isozymes contain two zinc finger
motifs, designated “C1a” and “C1b” domains, which
constitute the recognition modules for the second messenger diacylglycerol
(DAG) or the phorbol esters. However, the individual contributions of these
tandem C1 domains to PKC function and, reciprocally, the influence of protein
context on their function remain uncertain. In the present study, we prepared
PKCδ constructs in which the individual C1a and C1b domains were
deleted, swapped, or substituted for one another to explore these issues. As
isolated fragments, both the δC1a and δC1b domains potently bound
phorbol esters, but the binding of [3H]phorbol 12,13-dibutyrate
([3H]PDBu) by the δC1a domain depended much more on the
presence of phosphatidylserine than did that of the δC1b domain. In
intact PKCδ, the δC1b domain played the dominant role in
[3H]PDBu binding, membrane translocation, and down-regulation. A
contribution from the δC1a domain was nonetheless evident, as shown by
retention of [3H]PDBu binding at reduced affinity, by increased
[3H]PDBu affinity upon expression of a second δC1a domain
substituting for the δC1b domain, and by loss of persistent plasma
membrane translocation for PKCδ expressing only the δC1b domain,
but its contribution was less than predicted from the activity of the isolated
domain. Switching the position of the δC1b domain to the normal position
of the δC1a domain (or vice versa) had no apparent effect on the
response to phorbol esters, suggesting that the specific position of the C1
domain within PKCδ was not the primary determinant of its activity.One of the essential steps for protein kinase C
(PKC)2 activation is
its translocation from the cytosol to the membranes. For conventional
(α, βI, βII, and γ) and novel (δ, ε, η,
and θ) PKCs, this translocation is driven by interaction with the
lipophilic second messenger sn-1,2-diacylglycerol (DAG), generated
from phosphatidylinositol 4,5-bisphosphate upon the activation of
receptor-coupled phospholipase C or indirectly from phosphatidylcholine via
phospholipase D (1). A pair of
zinc finger structures in the regulatory domain of the PKCs, the
“C1” domains, are responsible for the recognition of the DAG
signal. The DAG-C1 domain-membrane interaction is coupled to a conformational
change in PKC, both causing the release of the pseudosubstrate domain from the
catalytic site to activate the enzyme and triggering the translocation to the
membrane (2). By regulating
access to substrates, PKC translocation complements the intrinsic enzymatic
specificity of PKC to determine its substrate profile.The C1 domain is a highly conserved cysteine-rich motif (∼50 amino
acids), which was first identified in PKC as the interaction site for DAG or
phorbol esters (3). It
possesses a globular structure with a hydrophilic binding cleft at one end
surrounded by hydrophobic residues. Binding of DAG or phorbol esters to the C1
domain caps the hydrophilic cleft and forms a continuous hydrophobic surface
favoring the interaction or penetration of the C1 domain into the membrane
(4). In addition to the novel
and classic PKCs, six other families of proteins have also been identified,
some of whose members possess DAG/phorbol ester-responsive C1 domains. These
are the protein kinase D (5),
the chimaerin (6), the munc-13
(7), the RasGRP (guanyl
nucleotide exchange factors for Ras and Rap1)
(8), the DAG kinase
(9), and the recently
characterized MRCK (myotonic dystrophy kinase-related
Cdc42-binding kinase) families
(10). Of these C1
domain-containing proteins, the PKCs have been studied most extensively and
are important therapeutic targets
(11). Among the drug
candidates in clinical trials that target PKC, a number such as bryostatin 1
and PEP005 are directed at the C1 domains of PKC rather than at its catalytic
site.Both the classic and novel PKCs contain in their N-terminal regulatory
region tandem C1 domains, C1a and C1b, which bind DAG/phorbol ester
(12). Multiple studies have
sought to define the respective roles of these two C1 domains in PKC
regulation, but the issue remains unclear. Initial in vitro binding
measurements with conventional PKCs suggested that 1 mol of phorbol ester
bound per mole of PKC
(13-15).
On the other hand, Stubbs et al., using a fluorescent phorbol ester
analog, reported that PKCα bound two ligands per PKC
(16). Further, site-directed
mutagenesis of the C1a and C1b domains of intact PKCα indicated that the
C1a and C1b domains played equivalent roles for membrane translocation in
response to phorbol 12-myristate 13-acetate (PMA) and (-)octylindolactam V
(17). Likewise, deletion
studies indicated that the C1a and C1b domains of PKCγ bound PDBu
equally with high potency (3,
18). Using a functional assay
with PKCα expression in yeast, Shieh et al.
(19) deleted individual C1
domains and reported that C1a and C1b were both functional and equivalent upon
stimulation by PMA, with either deletion causing a similar reduction in
potency of response, whereas for mezerein the response depended essentially on
the C1a domain, with much weaker response if only the C1b domain was present.
Using isolated C1 domains, Irie et al.
(20) suggested that the C1a
domain of PKCα but not those of PKCβ or PKCγ bound
[3H]PDBu preferentially; different ligands showed a generally
similar pattern but with different extents of selectivity. Using synthesized
dimeric bisphorbols, Newton''s group reported
(21) that, although both C1
domains of PKCβII are oriented for potential membrane interaction, only
one C1 domain bound ligand in a physiological context.In the case of novel PKCs, many studies have been performed on PKCδ
to study the equivalency of the twin C1 domains. The P11G point mutation of
the C1a domain, which caused a 300-fold loss of binding potency in the
isolated domain (22), had
little effect on the phorbol ester-dependent translocation of PKCδ in
NIH3T3 cells, whereas the same mutation of the C1b caused a 20-fold shift in
phorbol ester potency for inducing translocation, suggesting a major role of
the C1b domain for phorbol ester binding
(23). A secondary role for the
C1a domain was suggested, however, because mutation in the C1a domain as well
as the C1b domain caused a further 7-fold shift in potency. Using the same
mutations in the C1a and C1b domains, Bögi et al.
(24) found that the binding
selectivity for the C1a and C1b domains of PKCδ appeared to be
ligand-dependent. Whereas PMA and the indole alkaloids indolactam and
octylindolactam were selectively dependent on the C1b domain, selectivity was
not observed for mezerein, the 12-deoxyphorbol 13-monoesters prostratin and
12-deoxyphorbol 13-phenylacetate, and the macrocyclic lactone bryostatin 1
(24). In in vitro
studies using isolated C1a and C1b domains of PKCδ, Cho''s group
(25) described that the two C1
domains had opposite affinities for DAG and phorbol ester; i.e. the
C1a domain showed high affinity for DAG and the C1b domain showed high
affinity for phorbol ester. No such difference in selectivity was observed by
Irie et al. (20).PKC has emerged as a promising therapeutic target both for cancer and for
other conditions, such as diabetic retinopathy or macular degeneration
(26-30).
Kinase inhibitors represent one promising approach for targeting PKC, and
enzastaurin, an inhibitor with moderate selectivity for PKCβ relative to
other PKC isoforms (but still with activity on some other non-PKC kinases) is
currently in multiple clinical trials. An alternative strategy for drug
development has been to target the regulatory C1 domains of PKC. Strong proof
of principle for this approach is provided by multiple natural products,
e.g. bryostatin 1 and PEP005, which are likewise in clinical trials
and which are directed at the C1 domains. A potential advantage of this
approach is the lesser number of homologous targets, <30 DAG-sensitive C1
domains compared with over 500 kinases, as well as further opportunities for
specificity provided by the diversity of lipid environments, which form a
half-site for ligand binding to the C1 domain. Because different PKC isoforms
may induce antagonistic activities, inhibition of one isoform may be
functionally equivalent to activation of an antagonistic isoform
(31).Along with the benzolactams
(20,
32), the DAG lactones have
provided a powerful synthetic platform for manipulating ligand: C1 domain
interactions (31). For
example, the DAG lactone derivative 130C037 displayed marked selectivity among
the recombinant C1a and C1b domains of PKCα and PKCδ as well as
substantial selectivity for RasGRP relative to PKCα
(33). Likewise, we have shown
that a modified DAG lactone (dioxolanones) can afford an additional point of
contact in ligand binding to the C1b domain of PKCδ
(34). Such studies provide
clear examples that ligand-C1 domain interactions can be manipulated to yield
novel patterns of recognition. Further selectivity might be gained with
bivalent compounds, exploiting the spacing and individual characteristics of
the C1a and C1b domains (35).
A better understanding of the differential roles of the two C1 domains in PKC
regulation is critical for the rational development of such compounds. In this
study, by molecularly manipulating the C1a or C1b domains in intact
PKCδ, we find that both the C1a and C1b domains play important roles in
PKCδ regulation. The C1b domain is predominant for ligand binding and
for membrane translocation of the whole PKCδ molecule. The C1a domain of
intact PKCδ plays only a secondary role in ligand binding but stabilizes
the PKCδ molecule at the plasma membrane for downstream signaling. In
addition, we show that the effect of the individual C1 domains of PKCδ
does not critically depend on their position within the regulatory domain. 相似文献
12.
V. G. Ladygin 《Applied Biochemistry and Microbiology》2014,50(6):644-651
The composition of the carotenes and xanthophylls of Chlamydomonas reinhardtii Dang. C-41, a mutant of a unicellular green alga and a superproducer of ζ-carotene, was studied. The light-harvesting complexes and a complex of the PS-II reaction center were established to be disrupted in the C-41 mutant. However, the mutant retained a high (up to 46%) photosynthetic activity and the capacity to accumulate chlorophylls and carotenoids (up to 50%). The composition of carotenes was studied, and it was shown that, in contrast to wild-type K(+) cells, which accumulate up to 95% of β-carotene and 5% α-carotene, cells of the C-41 mutant contained 43% β-carotene, 19% β-zeacarotene, and 38% ζ-carotene. The high level of C-41 mutant biomass accumulation made it possible to recommend the mutant as a superproducer of ζ-carotene in phytobiotechnology. 相似文献
13.
Omar Ramadan Yongxia Qu Raj Wadgaonkar Ghayath Baroudi Eddy Karnabi Mohamed Chahine Mohamed Boutjdir 《The Journal of biological chemistry》2009,284(8):5042-5049
The novel α1D L-type Ca2+ channel is expressed
in supraventricular tissue and has been implicated in the pacemaker activity
of the heart and in atrial fibrillation. We recently demonstrated that PKA
activation led to increased α1D Ca2+ channel
activity in tsA201 cells by phosphorylation of the channel protein. Here we
sought to identify the phosphorylated PKA consensus sites on the
α1 subunit of the α1D Ca2+
channel by generating GST fusion proteins of the intracellular loops, N
terminus, proximal and distal C termini of the α1 subunit of
α1D Ca2+ channel. An in vitro PKA kinase
assay was performed for the GST fusion proteins, and their phosphorylation was
assessed by Western blotting using either anti-PKA substrate or
anti-phosphoserine antibodies. Western blotting showed that the N terminus and
C terminus were phosphorylated. Serines 1743 and 1816, two PKA consensus
sites, were phosphorylated by PKA and identified by mass spectrometry. Site
directed mutagenesis and patch clamp studies revealed that serines 1743 and
1816 were major functional PKA consensus sites. Altogether, biochemical and
functional data revealed that serines 1743 and 1816 are major functional PKA
consensus sites on the α1 subunit of α1D
Ca2+ channel. These novel findings provide new insights into the
autonomic regulation of the α1D Ca2+ channel in
the heart.L-type Ca2+ channels are essential for the generation of normal
cardiac rhythm, for induction of rhythm propagation through the
atrioventricular node and for the contraction of the atrial and ventricular
muscles
(1–5).
L-type Ca2+ channel is a multisubunit complex including
α1, β and α2/δ subunits
(5–7).
The α1 subunit contains the voltage sensor, the selectivity
filter, the ion conduction pore, and the binding sites for all known
Ca2+ channel blockers
(6–9).
While α1C Ca2+ channel is expressed in the atria
and ventricles of the heart
(10–13),
expression of α1D Ca2+ channel is restricted to
the sinoatrial (SA)2
and atrioventricular (AV) nodes, as well as in the atria, but not in the adult
ventricles (2,
3,
10).Only recently it has been realized that α1D along with
α1C Ca2+ channels contribute to L-type
Ca2+ current (ICa-L) and they both play important but
unique roles in the physiology/pathophysiology of the heart
(6–9).
Compared with α1C, α1D L-type
Ca2+ channel activates at a more negative voltage range and shows
slower current inactivation during depolarization
(14,
15). These properties may
allow α1D Ca2+ channel to play critical roles in
SA and AV nodes function. Indeed, α1D Ca2+ channel
knock-out mice exhibit significant SA dysfunction and various degrees of AV
block (12,
16–19).The modulation of α1C Ca2+ channel by
cAMP-dependent PKA phosphorylation has been extensively studied, and the C
terminus of α1 was identified as the site of the modulation
(20–22).
Our group was the first to report that 8-bromo-cAMP (8-Br-cAMP), a
membrane-permeable cAMP analog, increased α1D Ca2+
channel activity using patch clamp studies
(2). However, very little is
known about potential PKA phosphorylation consensus motifs on the
α1D Ca2+ channel. We therefore hypothesized that
the C terminus of the α1 subunit of the α1D
Ca2+ channel mediates its modulation by cAMP-dependent PKA
pathway. 相似文献
14.
15.
16.
17.
18.
Yeong-Su Kim Nam-Hee Kim Soo-Jin Yeom Seon-Won Kim Deok-Kun Oh 《The Journal of biological chemistry》2009,284(23):15781-15793
Codon optimization was used to synthesize the blh gene from the uncultured marine bacterium 66A03 for expression in Escherichia coli. The expressed enzyme cleaved β-carotene at its central double bond (15,15′) to yield two molecules of all-trans-retinal. The molecular mass of the native purified enzyme was ∼64 kDa as a dimer of 32-kDa subunits. The Km, kcat, and kcat/Km values for β-carotene as substrate were 37 μm, 3.6 min−1, and 97 mm−1 min−1, respectively. The enzyme exhibited the highest activity for β-carotene, followed by β-cryptoxanthin, β-apo-4′-carotenal, α-carotene, and γ-carotene in decreasing order, but not for β-apo-8′-carotenal, β-apo-12′-carotenal, lutein, zeaxanthin, or lycopene, suggesting that the presence of one unsubstituted β-ionone ring in a substrate with a molecular weight greater than C35 seems to be essential for enzyme activity. The oxygen atom of retinal originated not from water but from molecular oxygen, suggesting that the enzyme was a β-carotene 15,15′-dioxygenase. Although the Blh protein and β-carotene 15,15′-monooxygenases catalyzed the same biochemical reaction, the Blh protein was unrelated to the mammalian β-carotene 15,15′-monooxygenases as assessed by their different properties, including DNA and amino acid sequences, molecular weight, form of association, reaction mechanism, kinetic properties, and substrate specificity. This is the first report of in vitro characterization of a bacterial β-carotene-cleaving enzyme.Vitamin A (retinol) is a fat-soluble vitamin and important for human health. In vivo, the cleavage of β-carotene to retinal is an important step of vitamin A synthesis. The cleavage can proceed via two different biochemical pathways (1, 2). The major pathway is a central cleavage catalyzed by mammalian β-carotene 15,15′-monooxygenases (EC 1.14.99.36). β-Carotene is cleaved by the enzyme symmetrically into two molecules of all-trans-retinal, and retinal is then converted to vitamin A in vivo (3–5). The second pathway is an eccentric cleavage that occurs at double bonds other than the central 15,15′-double bond of β-carotene to produce β-apo-carotenals with different chain lengths, which are catalyzed by carotenoid oxygenases from mammals, plants, and cyanobacteria (6). These β-apo-carotenals are degraded to one molecule of retinal, which is subsequently converted to vitamin A in vivo (2).β-Carotene 15,15′-monooxygenase was first isolated as a cytosolic enzyme by identifying the product of β-carotene cleavage as retinal (7). The characterization of the enzyme and the reaction pathway from β-carotene to retinal were also investigated (4, 8). The enzyme activity has been found in mammalian intestinal mucosa, jejunum enterocytes, liver, lung, kidney, and brain (5, 9, 10). Molecular cloning, expression, and characterization of β-carotene 15,15′-monooxygenase have been reported from various species, including chickens (11), fruit flies (12), humans (13), mice (14), and zebra fishes (15).Other proteins thought to convert β-carotene to retinal include bacterioopsin-related protein (Brp) and bacteriorhodopsin-related protein-like homolog protein (Blh) (16). Brp protein is expressed from the bop gene cluster, which encodes the structural protein bacterioopsin, consisting of at least three genes as follows: bop (bacterioopsin), brp (bacteriorhodopsin-related protein), and bat (bacterioopsin activator) (17). brp genes were reported in Haloarcula marismortui (18), Halobacterium sp. NRC-1 (19), Halobacterium halobium (17), Haloquadratum walsbyi, and Salinibacter ruber (20). Blh protein is expressed from the proteorhodopsin gene cluster, which contains proteorhodopsin, crtE (geranylgeranyl-diphosphate synthase), crtI (phytoene dehydrogenase), crtB (phytoene synthase), crtY (lycopene cyclase), idi (isopentenyl diphosphate isomerase), and blh gene (21). Sources of blh genes were previously reported in Halobacterium sp. NRC-1 (19), Haloarcula marismortui (18), Halobacterium salinarum (22), uncultured marine bacterium 66A03 (16), and uncultured marine bacterium HF10 49E08 (21). β-Carotene biosynthetic genes crtE, crtB, crtI, crtY, ispA, and idi encode the enzymes necessary for the synthesis of β-carotene from isopentenyl diphosphate, and the Idi, IspA, CrtE, CrtB, CrtI, and CrtY proteins have been characterized in vitro (23–28). Blh protein has been proposed to catalyze or regulate the conversion of β-carotene to retinal (29, 30), but there is no direct proof of the enzymatic activity.In this study, we used codon optimization to synthesize the blh gene from the uncultured marine bacterium 66A03 for expression in Escherichia coli, and we performed a detailed biochemical and enzymological characterization of the expressed Blh protein. In addition, the properties of the enzyme were compared with those of mammalian β-carotene 15,15′-monooxygenases. 相似文献
19.
20.
Uptake of HCO3? and CO2 in Cells and Chloroplasts from the Microalgae Chlamydomonas reinhardtii and Dunaliella tertiolecta 下载免费PDF全文
Gabi Amoroso Dieter Sültemeyer Christoph Thyssen Heinrich P. Fock 《Plant physiology》1998,116(1):193-201
Mass-spectrometric disequilibrium analysis was applied to investigate CO2 uptake and HCO3− transport in cells and chloroplasts of the microalgae Dunaliella tertiolecta and Chlamydomonas reinhardtii, which were grown in air enriched with 5% (v/v) CO2 (high-Ci cells) or in ambient air (low-Ci cells). High- and low-Ci cells of both species had the capacity to transport CO2 and HCO3−, with maximum rates being largely unaffected by the growth conditions. In high- and low-Ci cells of D. tertiolecta, HCO3− was the dominant inorganic C species taken up, whereas HCO3− and CO2 were used at similar rates by C. reinhardtii. The apparent affinities of HCO3− transport and CO2 uptake increased 3- to 9-fold in both species upon acclimation to air. Photosynthetically active chloroplasts isolated from both species were able to transport CO2 and HCO3−. For chloroplasts from C. reinhardtii, the concentrations of HCO3− and CO2 required for half-maximal activity declined from 446 to 33 μm and 6.8 to 0.6 μm, respectively, after acclimation of the parent cells to air; the corresponding values for chloroplasts from D. tertiolecta decreased from 203 to 58 μm and 5.8 to 0.5 μm, respectively. These results indicate the presence of inducible high-affinity HCO3− and CO2 transporters at the chloroplast envelope membrane. 相似文献