Departments of Obstetrics and Gynaecology and Biochemistry, Canadian Institutes of Health Research Group in Fetal and Neonatal Health and Development, The University of Western Ontario, London, Ontario, Canada N6A 5A5
![]() |
ABSTRACT |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
The lung contains two distinct forms of phosphatidic acid phosphatase (PAP). PAP1 is a cytosolic enzyme that is activated through fatty acid-induced translocation to the endoplasmic reticulum, where it converts phosphatidic acid (PA) to diacylglycerol (DAG) for the biosynthesis of phospholipids and neutral lipids. PAP1 is Mg2+ dependent and sulfhydryl reagent sensitive. PAP2 is a six-transmembrane-domain integral protein localized to the plasma membrane. Because PAP2 degrades sphingosine-1-phosphate (S1P) and ceramide-1-phosphate in addition to PA and lyso-PA, it has been renamed lipid phosphate phosphohydrolase (LPP). LPP is Mg2+ independent and sulfhydryl reagent insensitive. This review describes LPP isoforms found in the lung and their location in signaling platforms (rafts/caveolae). Pulmonary LPPs likely function in the phospholipase D pathway, thereby controlling surfactant secretion. Through lowering the levels of lyso-PA and S1P, which serve as agonists for endothelial differentiation gene receptors, LPPs regulate cell division, differentiation, apoptosis, and mobility. LPP activity could also influence transdifferentiation of alveolar type II to type I cells. It is considered likely that these lipid phosphohydrolases have critical roles in lung morphogenesis and in acute lung injury and repair.
epidermal growth factor receptors; caveolae; lysophospholipids; phospholipase D; surfactant secretion
![]() |
INTRODUCTION |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
SINCE 1899, when Overton (186) reported that the diffusion of chemical compounds into cells varied directly with their solubility in oil compared with water, it has been recognized that lipids are major constituents of plasma membranes. Subsequent studies by Gorter and Grendel (77) and Danielli and Davson (47) in the 1920s and 1930s concluded that the limiting membrane at the cell surface was a bimolecular (phospho)lipid leaflet (see Ref. 286 for further details). Numerous investigations, in particular electron microscopic examinations of cellular and reconstituted membranes (207, 240), led to the conclusion that intracellular and extracellular biological membranes share similar structural properties. Further studies on membrane properties eventually led to the presently accepted fluid-mosaic model (229).
More recently, biochemists and cell biologists have focused considerable attention on mechanisms by which information can be transmitted across the hydrophobic barriers surrounding biological cells. Interestingly, it has become apparent that the phospholipid constituents of cells provide substrates for first and second messengers and that phospholipases, originally thought of only as degradative enzymes, play important roles in signaling. For example, early observations revealed that phospholipase A2 (PLA2), by releasing the fatty acids from the sn-2 position of certain phospholipids, provides arachidonic acid for eicosanoid synthesis (51, 72, 133). Hydrolysis of phosphatidylinositol-4,5 bis phosphate (PIP2) by phosphatidylinositol-specific phospholipase C (PI-PLC) provides both diacylglycerol (DAG) and inositol tris phosphate (IP3) for protein kinase C (PKC) activation and intracellular calcium mobilization (202). In addition, it has become evident that phospholipids themselves can act as first and second messengers. PLA2 hydrolysis of 1-alkyl,2-arachidonoyl phosphatidylcholine not only generates substrate for eicosanoid formation but provides the lyso-precursor for 1-alkyl,2-acetyl phosphatidylcholine, also known as platelet-activating factor (195). Also, PIP2 leads to intracellular messages, not only through degradation by PI-PLC but also via phosphatidylinositol 3-kinase activity, resulting in PI 3,4,5-trisphosphate (PIP3), an important activator of protein kinase B (21). More germane to the present review, lysophosphatidic acid (LPA) and sphingosine-1-phosphate (S1P) are important serum-derived extracellular first messengers, with potent physiological effects.
The present article focuses on phosphatidic acid (PA) phosphatase (PAP), which hydrolyzes 1,2-diacyl,sn-glycerol-3-phosphate to DAG and inorganic phosphate. Originally considered only for its role in glycerolipid biosynthesis, it has recently been demonstrated that this phosphohydrolase plays an important role as a modulator of cell signaling in relation to cell mobility, differentiation, growth, and survival (24, 25, 31, 120, 260).
Current evidence indicates at least two distinct classes of PAP in mammalian cells. PAP1 is a soluble cytosolic protein, which, under the influence of free fatty acids and perhaps other signals, can translocate to the endoplasmic reticulum (ER) and possibly other intracellular organelles, where it converts newly biosynthesized phosphatidate to DAG. DAG generated by PAP1 is utilized for the formation of phosphatidylcholine (PC), phosphatidylethanolamine (PE), and triacylglycerol (TAG). Although so far intransigent to purification, PAP1 has been shown to be Mg2+ dependent and heat and sulphhydryl reagent sensitive.
The second class of phosphatidate phosphatase (PAP2) comprises a family
of six-transmembrane-domain glycosylated proteins localized in the
plasma membrane. Unlike PAP1, which is highly specific for the
Mg2+ salts of PA and LPA, PAP2 degrades a number of lipid
phosphates, including S1P, ceramide-1-phosphate (C1P), as well as PA
and LPA, does not require specific ions, and is heat and sulphhydryl
reagent resistant. It has consequently been suggested that PAP2 be
designated lipid phosphate phosphatase (LPP) (25, 107,
260). Although not yet accepted by scientific reference bases
such as PubMed, which still uses phosphatidic acid phosphatase, this
recently suggested nomenclature emphasizes the molecular distinction
between PAP1 and PAP2, thereby eliminating past confusion regarding
these enzymes. In addition, the new enzyme system (Table
1) greatly simplifies current literature
regarding the variously named LPP isoforms that have only recently been
discovered. Therefore, this rationalizes nomenclature and will be used
here with the previous term included in parentheses as the different
isoforms are introduced. The present review will summarize our present
understanding of PAP1 and LPP and their physiological functions, with
particular emphasis on pulmonary phosphatidate phosphohydrolase. For
brevity, liberal use will be made of earlier reviews.
|
![]() |
PAP1 AND LPP ENZYMATIC ACTIVITIES |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Phosphatidic acid phosphohydrolase was initially described in
plants by Morris Kates in 1955 (123). Subsequently, the
dephosphorylation of PA was reported in a number of mammalian tissues
(235). Original interest in this activity was related to
its role in phospholipid and neutral glycerolipid biosynthesis. PA is
situated at an important branch point in the Kennedy de novo pathway
(124, 125), leading to the formation of either anionic
(acidic) phospholipids or zwitterionic and neutral lipids (Fig.
1). DAG generated by PAP1 acts as a
substrate for biogenesis of PC, PE, and TAG.
|
After its discovery in plants (123), phosphatidate phosphohydrolase in mammals was initially characterized in the early 1960s in the brain (2), intestine (111), kidney (41), erythrocytes (92), and liver (225, 273). These investigations, which used relatively high concentrations of phosphatidate as aqueous emulsions in the absence of Mg2+, found that this activity (i.e., LPP) was associated with particulate fractions with little or no activity present in the high-speed cytosol. Paradoxically, despite the high phosphohydrolase activity observed with microsomes and mitochondria, radioactivity from labeled glycerol-3-phosphate accumulated in PA (97, 233, 239, 244). Further investigations revealed the cytosolic fraction contained a factor that stimulated production of DAG and TAG from labeled phosphatidate. Although initially controversial, investigations with liver, intestine, adipose tissue, and lung eventually revealed that the cytosolic factor was soluble PAP1 (106, 112, 161, 234, 265; reviewed in Refs. 22, 23, 194).
It is instructive to consider the basis of the initial inability to identify phosphatidate phosphohydrolase in cell supernatants. In retrospect, a primary reason was that the chemical PA substrate used was generated by plant phospholipase D (PLD), which requires Ca2+. This divalent cation inhibits Mg2+-independent PAP1 but has a much smaller effect on LPP activity. Calcium phosphatidate is very stable, and until the development of the ion-exchange resin Chelex 100 (203), it was virtually impossible to remove this divalent cation completely from PA preparations. Calcium phosphatidate can form hexagonal II (HII) phase where the phosphate groups would be internalized away from the dispersing buffer (95), thereby limiting access to these groups for enzymatic hydrolysis. A second reason for the initial failure to detect Mg2+-dependent PAP1 activity is that LPP possesses higher Km and Vmax. As a result, with most tissues, Mg2+-dependent (i.e., PAP1) activity could not be detected, especially under conditions more or less optimized for LPP. In addition, many studies employed detergents such as Tween 20 and Triton X, which stimulate LPP but markedly inhibit the more sensitive PAP1 (263).
PAP1 activity was first recognized using radiolabeled PA biosynthetically generated on microsomal or mitochondrial membranes. The ability of the cytosolic PAP1 to hydrolyze biosynthetically generated, but not aqueously dispersed, PA was attributed to the need for substrate presentation in a membrane-associated form (33). This "natural" form, which persisted in autoclaved microsomes, was eventually found to be related to the presence of Mg2+ during generation of the membrane-associated substrate; i.e., the Mg2+ salt of PA was formed, and this was the true substrate for PAP1. Subsequently, it was shown that mixing PC with the radiolabeled PA markedly increased the PAP1 activity. Thus the requirement for a natural membrane-associated substrate could be achieved by preparing vesicles of mixed PA/PC and providing optimal Mg2+ (23, 194, 242, 264, 265).
Such chemically defined substrates permitted investigators to demonstrate that PAP1 was predominantly localized to the cytosol but translocated to the ER and perhaps other organelles under the influence of free fatty acids, acyl CoAs, and possibly other agents. Washing microsomes with high salt virtually abolished their ability to generate DAG, TAG, or PC from radioactive glycerol-3-phosphate. These biosynthetic capacities were restored by adding back the washings or partially purified PAP1 (265). These investigations showed that PAP1 was involved in glycerolipid biosynthesis. They also demonstrated that the Mg2+-independent, heat-stable LPP activity, which was not removed by washing, could not contribute significantly to the de novo pathway shown in Fig. 1. The biological role of the abundant LPP activities retained by microsomal fractions remained unclear until Brindley and colleagues (105, 258, 259) demonstrated localization of this Mg2+-independent, N-ethylmaleimide (NEM)-insensitive enzyme in hepatocyte plasma membranes, suggesting a potential role in signal transduction. This group also demonstrated that, in addition to PA and LPA, LPP was also capable of hydrolyzing S1P and C1P, which are potent signaling lipid phosphates.
![]() |
SUBSTRATES FOR LPP |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
This section will briefly discuss, with reference to helpful reviews, the enzymatic mechanisms responsible for producing lipid substrates for LPP. As indicated above, present evidence, albeit circumstantial, indicates LPP does not have access to LPA or PA generated during de novo glycerolipid synthesis via the Kennedy pathway (22, 194, 260) (Fig. 1). Neither S1P nor C1P is considered a biosynthetic intermediate. PA is generated on plasma membranes and other cellular membranes by PLD and DAG kinase in response to cytokines, hormones, neurotransmitters, and related agonists (61, 69, 113, 140, 250, 255). LPA and S1P are present in serum and therefore available to LPP on plasma membranes. C1P can also be generated on plasma membranes in response to certain signals (24, 196).
PLD.
PLD was initially discovered in plants by Hanahan and Chaikoff in 1947 (84). Early studies (279) demonstrated that
in addition to hydrolyzing phospholipids by introducing H+
and OH (i.e., H2O) across the phosphodiester
bond, PLD catalyzes a transphosphatidylation reaction in which a
short-chained alcohol replaces water to generate a new
phosphatidylalcohol instead of PA and releases the original alcohol,
which is choline in the case of PC.
|
DAG kinase. In addition to PLD, LPP could act on the PA produced by DAG kinase. A primary role for DAG kinase appears to be the termination of PKC activation by removing DAG generated by PI-PLC (250, 255, 261). The PA produced in this manner is used for formation of CDP-diacylglycerol and PI synthesis similar to the de novo pathway in Fig. 1. The resulting PI is subsequently sequentially phosphorylated to PIP2 (13-16, 53).
Nine DAG kinases have been identified that share a similar catalytic domain but show differences in primary structure, substrate specificity, and tissue distribution (250, 255). In addition to the catalytic domain, DAG kinases contain a number of other conserved motifs crucial for lipid-protein and protein-protein interactions. The occurrence of nine distinct DAG kinase isoforms (to date) in mammalian tissue implies regulatory differences. DAG kinases can be controlled through translocation to plasma and nuclear membranes, by calcium, and through phosphorylation. One isoform, DAG kinase-Lipid phosphate growth factors. It has long been apparent that in addition to peptide growth factors such as insulin-like growth factors, serum contains phospholipid growth factors, including LPA, S1P, and related compounds (3, 76, 89, 164, 199). LPA and S1P, present in serum at micromolar concentrations, are released by activated platelets, stimulated leukocytes, and cells in apoptosis that shed microvesicles from their plasma membranes. Calcium influx activates scramblase, a protein that accelerates "flip-flop" of phospholipids in a nonspecific manner that eliminates the membrane asymmetry inherent to healthy cells (17, 171, 200). The loss of membrane phospholipid asymmetry contributes to microvesicle shedding. Calcium also stimulates PLD activity leading to the formation of PA. In addition, conditions promoting platelet and leukocyte activation can also stimulate release of group II secretory PLA2. This phospholipase, which has a high affinity for acidic phospholipids such as PA (236), generates LPA and other lysophospholipids. The manner in which PA generated by PLD on the inner leaflet of the plasma membrane becomes available for PLA2 action on the outer leaflet is not known, but scramblase may have a role.
LPA can also be formed through PLD hydrolysis of lysophospholipids generated after activation of cytosolic group V PLA2 (73, 274). A recently discovered lyso-PLD present in serum could function in this process (247). This mechanism for LPA formation, which has been demonstrated after ![]() |
CDNAS FOR LPPS |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Using NH2-terminal amino acid sequence from a 35-kDa
LPP purified from porcine thymus, Kai et al. (118)
observed high conservation with internal sequence from Hic 53, a mouse
partial cDNA originally identified as a
H2O2-inducible gene (57). The
predicted amino acid sequence of mouse LPP1 demonstrated 34% identity
with Drosophila Wunen and 48% identity with rat
Dri 42. Wunen appears involved in guiding germ cell
migration from the developing gut toward overlying mesenchyme early in
Drosophila embryonic development (283). A
second LPP cDNA, Dri 42 (LPP3/PAP2b), was identified as a novel gene
upregulated during the differentiation of rat intestinal mucosa
epithelium (12). Subsequently, Kai et al. (117) cloned the cDNAs for human LPP1 and LPP3 cDNAs from
Hep G2 cells, and Hooks et al. (93) reported the cloning
of LPP2 (PAP2c) from expressed sequence tags. Roberts et al.
(206) cloned human LPP1, LPP2, and LPP3 and showed these
clones expressed active enzyme in human embryonic kidney (HEK 293)
cells. The human cDNAs showed 40-60% sequence identity to each
other but arise from distinct genes, with the human expressing LPP1
being located on chromosome 5 (5q11), the gene for LPP2 on chromosome
19 (19p13), and the gene for LPP3 on chromosome 1 (1pter-p22.1).
Additionally, Leung et al. (135) reported an apparent
alternate splice variant, LPP1a (PAP2a-1), of human LPP1 from a
human lung cDNA library.
Apparent differences in substrate specificity between the Mg2+-independent NEM-resistant phosphatidate phosphohydrolase activities in rat lung (174) and liver (259) led Nanjundan and Possmayer (175) to apply reverse transcriptase-polymerase chain reaction (RT-PCR) to clone the pulmonary isoforms from rat lung and type II cell RNA. cDNAs were obtained for LPP1, three apparent alternate splice variants of LPP1 (LPP1a, LPP1b, and LPP1c), and LPP3. These rat LPP cDNAs exhibit ~50% identity at the amino acid level. Attempts to demonstrate LPP2 in rat lung using PCR primers based on the human sequences were unsuccessful. Human LPP2 expression appears to be limited to the brain, pancreas, and placenta. It has recently been found that in the mouse, LPP2 is expressed in the lung, liver, and kidney (282). As a result LPP2 distribution in rat lung must be re-evaluated.
LPP1a, LPP1b, and LPP1c appear to be splice variants of LPP1. An
inspection of the cDNAs for LPP1 and these variants suggests four
regions: I, IIA, IIB, and III. As illustrated in Fig.
3, all four cDNAs contain regions
I and III. LPP1 cDNA, containing regions I,
IIB, and III, and LPP1a cDNA, containing regions
I, IIA, and III, encode for predicted proteins of 282 and 283 amino acids, respectively. The predicted proteins would contain
six transmembrane domains with the NH2 and COOH termini on
the cytoplasmic sides of plasma and other membranes and a consensus
N-linked glycosylation site at positions 142 (LPP1) and 143 (LPP1a) of
the second extracellular loop. LPP3 cDNA codes for a protein of 308 amino acids with a consensus N-linked glycosylation site at
Asn168. Rat lung LPP1 and LPP3 cDNAs are virtually
identical to the previously reported cDNAs for LPP1 in rat liver and
LPP3 in rat intestine (Dri 42), and rat lung LPP1a shows 80% amino
acid sequence identity to the human isoform. LPP1b and LPP1c are novel
isoforms that predict truncated forms of phosphatidate
phosphohydrolase.
|
LPPs contain a novel conserved phosphatase sequence motif,
K(X)6 RP- (X12-54)-PSGH-
(X31-54)-SR(X)5 H(X)3D, that is shared among several yeast lipid phosphatases, some mammalian phosphatases, including mammalian glucose-6-phosphatase,
chloroperoxidases, and some bacterial nonspecific acid phosphatases
(Fig. 4). Stukey and Carman
(241) have proposed a model that includes 1)
nucleophilic attack of the substrate's phosphoryl group by the
histidine of active site domain 3 and production of a
phosphoenzyme catalytic intermediate; 2) involvement of the
conserved arginine residues of active site domains 1 and
3 in hydrogen bonding to equatorial phosphoryl oxygens; and
3) participation of the histidine of active site
domain 2 in protonation of the substrate leaving group.
Active site domains are indicated such that domains 1 and
2 are located on the second extramembrane loop, whereas the
third active site region is present on the third extracellular loop of
the protein (Fig. 4). Zhang et al. (284) demonstrated that
mutating Lys120, Arg127, Pro128,
Ser169, His171, Arg217, or
His223 from mouse LPP1 results in a 95% or greater loss of
activity, whereas altering Gly170 reduces activity by
~66%. Altering the putative glycosylation site Asn142 to
Gln resulted in a ~4-kDa reduction in molecular mass. The nonglycosylated product remained enzymatically active as occurs with
enzymatic removal of the LPP glycan.
|
As indicated above, LPP1b and LPP1c code for truncated proteins. The LPP1b isoform did not contain either cDNA region IIA or IIB (Fig. 3), except for the insertion of a single G nucleotide at the I/III boundary. This insertion resulted in a frame shift leading to early termination of the predicted protein at 30 amino acids. The predicted protein, if formed, would possess an NH2-terminal extramembrane domain but would possess only part of the first transmembrane domain and differ in 10 of the last 11 amino acids at the predicted COOH terminus. The cDNA for LPP1c contains regions IIA and IIB, except for the deletion of a "G" nucleotide at the IIA-IIB boundary. This deletion also results in a frame shift, leading to early termination at 76 amino acids. Neither LPP1b nor LPP1c would possess an active site.
It must be stressed that whether proteins corresponding to LPP1b or LPP1c nucleotide sequences are produced in the lung or any other tissue remains unknown. LPP1c mRNA is rare, but LPP1b mRNA is relatively abundant. RNAase protection assays reveal that in the rat lung, the mRNA for LPP1b is ~25% of the level of LPP1, whereas LPP1a mRNA is only 10% of LPP1 mRNA levels (L. Zhao and F. Possmayer, unpublished results). RT-PCR studies have shown that LPP1b is relatively high in fetal lung but declines after birth, whereas LPP1a mRNA is low before birth but increases toward adulthood. RT-PCR, using primers designed to distinguish between LPP1, LPP1a, and LPP1b, shows LPP1 was present in rat liver, intestine, kidney, spleen, uterus, heart, and brain, as well as in the lung (175). LPP1b mRNA was particularly high in the lung, brain, intestine, and heart, whereas LPP1a was relatively high in the kidney, intestine, spleen, lung, and liver.
The observed variations in LPP1b expression relative to LPP1 and LPP1a
are consistent with biological regulation, but at present physiological
relevance remains unclear. It is noteworthy that precedent exists for
mRNAs coding for truncated forms of PLD, DAG kinase, and PLC-,
which, as in the case of LPP1b, would lack catalytic functions. For
example, Steed et al. (238) have reported PLD2c, a variant
of active PLD2a, contains a 56-bp insert resulting in premature
termination during RNA translation. Expression of the human PLD2c
variant mRNA relative to PLD2a is high in the liver and heart but low
in the brain and skeletal muscle. In addition, splice variants of human
PLD1 and PLD1b were found that possess 114-bp deletions resulting in
loss of an essential HKD transphosphatidylation motif, resulting in an
inactive enzyme.
A second potentially important example of isoform inactivation through
alternate splicing involves DAG kinase. As indicated earlier, DAG
kinase functions in the regeneration loop of the PI signaling cycle,
where it produces PA for PI synthesis via CDP-DAG but can also
downregulate PKC activity by removing DAG. The human retina
demonstrates high expression of a phosphatidylserine-dependent DAG
kinase- isoform (116). A low level of this full-length
active DAG kinase-
is also expressed in brain. Most other human
tissues express very low levels of the mRNA for the active DAG
kinase-
but express varying amounts of an mRNA coding for a
catalytically inactive form of this enzyme truncated by 25 amino acids
within the catalytic domain. Kai et al. (116) have
suggested that this represents a biological mechanism for
downregulating enzymatic expression of this protein at the mRNA
splicing level.
A third example of enzymatic downregulation occurs with an alternate
splice form of PKC, designated PKC-III, which is expressed in rat
testes (253). PKC-
III possesses an 83-nucleotide insert within the caspase-3 recognition domain, which produces premature truncation such that the protein retains regulatory segments but lacks
the catalytic domain. The parental PKC-
I isoform is cytosolic and
translocates to the plasma membrane upon phorbol ester stimulation. The
catalytic domain-deficient PKC-
III is predominantly localized on
plasma membranes and is only slightly more directed from cytosol to
peripheral membranes by phorbol esters.
These three precedents all involve phosphatidate metabolism. PA, a
substrate for plasma membrane LPP, is generated by DAG kinase and PLD,
whereas the LPP product DAG activates conventional and nonconventional
PKCs (50). It is possible that the LPP1b isoform could act
to control LPP activity by sequestering activation and/or inactivation
signals, as has been suggested by Ueyama et al. (253) for
PKC-. It is also conceivable that truncation via RNA splicing could
act as a physiological mechanism for reducing LPP1 and/or LPP1a isoform
mRNA expression in lung and other tissues. Attempts at overexpressing
LPP1, LPP1a, LPP1b, and LPP3 in mouse lung epithelial (MLE) cells using
the pTracer-CMV2 expression vector were unsuccessful due to cell death,
possibly by apoptosis (M. Nanjundan, R. White, A. Brickenden,
and F. Possmayer, unpublished results). This may have been
related to toxic effects resulting from the particularly strong
cytomegalovirus promoter during expression with this vector.
Alternatively, transient expression of LPP cDNAs could have led to
increases in bioactive sphingolipids implicated in programmed cell
death. However, transient expression of rat LPP1, LPP1a, and LPP3 in
HEK 293 cells resulted in increased LPP activity assayed with PA
without any suggestion of apoptosis (175), as has
been observed by others using LPP cDNAs from other species (118,
206). Transient expression of LPP1b in HEK 293 cells does not
affect endogenous LPP activity and had no apparent effect on cellular
viability (175). It is clear that further studies using
lung cells are needed to determine the physiological implications of
this LPP1 variant.
![]() |
RECEPTORS AND OTHER TARGETS/LIGANDS FOR LIPID PHOSPHATES |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
As indicated in SUBSTRATES FOR LPP, serum contains a number of lipid mediators, including LPA and S1P. Biological responses to these phospholipid growth factors include 1) calcium mobilization, 2) stimulation of cellular proliferation, 3) inhibition of apoptosis, 4) aggregation of platelets, 5) formation of stress fibers, 6) neurite contraction, 7) contraction of smooth muscle cells, 8) regulation of cell-cell interactions, 9) stimulation of cell mobility and 10) chemotaxis (3, 76, 165).
Endothelial differentiation gene receptors. The obvious physiological importance of the above indicated responses prompted attempts to clone putative receptors linking the extracellular presence of these lipid mediators to intracellular events. A number of difficulties were encountered related, in part, to the lipid nature of these agonists, their tendency to bind avidly to Ca2+, and even their variable presence in serum used for cell culture (3, 38, 89). Therefore, the demonstration by Chun et al. (38) that LPA was a natural ligand for ventricular zone-1 (vzg-1) receptor, so named because of its high expression in that particular neural proliferative zone in embryonic cortex, marked an important advance. Because it appeared likely that the LPA receptor vzg-1 was a member of the newly discovered endothelial differentiation gene (EDG) family of G protein-coupled receptors (i.e., EDG-2), this initial discovery facilitated identification of other lipid mediator ligands. The EDG-1 receptor was initially cloned by Hla and Maciag (90) in 1990 as a phorbol ester-inducible mRNA in human umbilical vein endothelial cells, which under the influence of serum, differentiate into capillary-like networks. S1P was identified as a potent mitogen arising from the phosphorylation of sphingosine (88). This lipid phosphate was later recognized as the polar agent in serum that promoted EDG-1-overexpressing HEK 293 cells to form tubular, capillary-like networks (88) and was subsequently found to act as a chemotactic factor in angiogenesis (59, 196). LPA and sphingosylphosphorylcholine were also shown to bind EDG-1 but with low affinity. Expression cloning studies using the serum response element and the jellyfish Ca2+-interacting protein apoaequorin confirmed LPA as a ligand for EDG-2 (3). Other members of this seven-transmembrane G protein-coupled family were soon found by various approaches. To date, three LPA high-affinity receptors (EDG-2, EDG-4, and EDG-7) and five S1P-responsive receptors (EDG-1, EDG-3, EDG-5, EDG-6, and EDG-8) have been identified. At a recent conference (Federation of American Societies for Experimental Biology Conference on Lysophospholipids; Tucson, AZ, June 9-14, 2001; Edward J. Goetzl, Timothy Hla, and Gabor Tigyi, organizers), there was a consensus to adopt a more logical nomenclature referring to these receptors in terms of their affinities, such that the receptors for LPA would be designated (with the old nomenclature after the slash) LPA1/EDG-2, LPA2/EDG-4, LPA3/EDG-7, whereas the receptors for S1P would be designated S1P1/EDG-1, S1P2/EDG-5, S1P3/EDG-3, S1P4/EDG-6, and S1P5/EDG-8 (Table 1). Adoption of the new nomenclature would conform with the International Union of Pharmacology recommendations and will alert readers to the chemical nature of the ligand. It appears likely that other receptors will be found for lipid phosphates (245).
Intracellular ligands for lipid phosphates. Although sphingosine kinase can be secreted, this enzyme is primarily present within cells (88). Sphingosine kinase activity on the inner aspect of the plasma membrane and possibly other cellular membranes would lead to the intracellular formation of S1P. The manner in which S1P accesses the outer aspect of the cell and interacts with S1P receptors is not known. Although S1P was originally thought of as an intracellular second messenger, a number of the actions attributed to intracellular S1P appear best explained through autocrine interactions with extracellular receptors (88, 196). However, as stressed by Pyne and Pyne (196), the location of S1P-mediated events is not always readily identifiable. The lack of or partial inhibition by pertussis toxin, which inhibits Gi/o but not Gs nor Gq, is inconclusive because neither the S1P2 nor the S1P4 receptor is affected by this toxin and their presence is difficult to exclude. The effects of the sphingosine analogs D,L-threo-dihydrosphingosine and N,N-dimethylsphingosine, which inhibit sphingosine kinase, are also ambiguous since these inhibitors can affect certain PKCs (196). Consequently, attributing potential second messenger roles for S1P must await identification of intracellular molecular targets and/or development of specific inhibitors for S1P receptors.
The manner in which PA interacts with cells is also controversial. PA is generated by PLD and DAG kinase, not only on the inner aspect of the plasma membrane but likely also on the cytosolic aspect of ER, vesicular, trans-Golgi, and nuclear membranes (18, 40, 113, 139). The manner in which PA becomes accessible to LPP-active sites on the outer face of the plasma membrane has not been ascertained. PA could also be formed on the luminal side of intracellular vesicles. After fusion with the plasma membrane, this PA would be accessible to LPP-active sites. In addition to LPP, PA could react with specific receptors, although direct evidence for such receptors is not available. PA can also be hydrolyzed to LPA and considerable evidence has accumulated implicating LPA as an extracellular messenger (18, 46, 113, 139, 156, 260). However, as in the case of S1P, caution is advised, and the evidence for a role for LPA or PA itself should be examined for each case. Certainly early reports implicating exogenous PA in intracellular Ca2+ release are clearly attributable to LPA contamination (165). Small amounts of LPA with PA likely explain effects on arachidonate release (172, 256) and PC-PLD activation (80, 257). These effects are attributable to LPA receptor interactions. In a number of cases, addition of bacterial PLD to intact or permeabilized cells elicits responses attributed to PA, although the possibility that DAG formation could be involved has not always been addressed (156). Nevertheless, current knowledge supports intracellular roles for PA. Apparent direct effects of PA have been noted on a large number of enzymes and enzyme systems, including PI-PLC (104, 108, 114, 248), cyclic nucleotide phosphodiesterase (150, 219), PI 4P-5 kinase (108, 166, 248), and a protein tyrosine phosphatase (249, 285). There are numerous reports of PA effects on protein kinases such as the PKC (138, 173, 280), protein kinase N, and p21cdcrac families (126, 129, 167, 267, 281). PA also appears to have inhibitory effects on Ras-GTPase-activating proteins (252) and Rho-GTP-dissociating isoforms (37; reviewed in Refs. 19, 115, 126, 155). The particular case of neutrophil respiratory heart NADPH oxidase complex, which generates H2O2 for the killing of pathogens, involves a 65-kDa PA-activated protein kinase that phosphorylates Ser/Thr residues on p22-phox and p47-phox (156). PA stimulation of NADPH oxidase is synergistic with DAG, but the neutral lipid target has not been identified. PMA does not replace DAG, and PKC does not appear to be involved (197). Evidence exists for a specific interaction between PA and the phosphorylation of Raf-1, a small GTPase of the Ras family (46). Phorbol esters induce prostaglandin synthetase transcription and translation in Madin Darby canine kidney cells (136). The most likely mechanism involves PKC- ![]() |
LPP IN SIGNALING PLATFORMS |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
The fluid-mosaic membrane model, as suggested by Singer and Nicolson (229), allowed for specialized membrane regions, such as caveolae (literally, "little caves"). Caveolae are nonclathrin-coated vesicular invaginations of the plasma membrane, often with a flask-like shape and possessing diameters of 50-100 nm (4, 139, 220). Large numbers of caveolae have been reported in endothelial cells, adipocytes, fibroblasts, and smooth muscle cells. Caveolae have not been detected in alveolar type II cells but are abundant in type I cells.
Caveolae are attached to the plasma membrane by a short neck, but they can also appear as flat pits. These latter structures could represent early stages of invagination (4). Caveolin, the principal structural component of caveolae, is a ~22-kDa scaffolding protein that acts as a marker for these membrane structures. Mammalian cells possess four forms of caveolin: caveolin-1a, caveolin-1b, caveolin-2, and caveolin-3 (220).
Interest in caveolae increased markedly with evidence that these are not fixed entities but dynamic structures that can bud from the plasma membrane and that such internalization is under the control of the molecular transport machinery responsible for vesicular budding, docking, and fusion (179, 223). It has further become evident that caveolae-like structures are involved in a number of cellular functions, including:
1) Transcellular transport, where caveolar vesicles move between two surfaces of the cell, transporting ions, small molecules, and low-molecular-weight macromolecules. This process is inhibited by NEM and cholesterol-binding agents such as filipin and appears to require GTP (4). Transmembrane transport has been studied extensively in endothelial cells but appears to function in placenta epithelial cells as well and may occur in alveolar type I cells (30, 178).
2) Potocytosis. This involves nonclathrin-coated pit transport of certain ions, low-molecular-weight molecules or macromolecules from the cell surface to the cytoplasm, to internal organelles such as the ER (4), and from one plasma membrane to another plasma membrane site (e.g., apical to basolateral). Folate transport, which involves a glycerolphosphoinositol (GPI)-anchored receptor, is a well-known example of potocytosis (160, 205), but Fe2+, Ca2+, alkaline phosphatase, and insulin are also transported in this manner (4, 85, 168). Caveolae-based endocytosis can be employed opportunistically by toxins, viruses, bacteria (227), and trypanosomes (63, 232) to gain access to the cell's interior.
3) Cholesterol transport. Caveolae appear involved in directing intracellular cholesterol transport (4, 64, 131, 141, 222). Caveolins bind cholesterol, and the scavenger receptor class B is a fatty acylated glycoprotein directed to caveolae, which mediates selective cholesterol uptake from the high-density lipoproteins. This contrasts with the low-density lipoproteins receptor-mediated transfer known to occur via clathrin-coated pits and vesicles (99).
4) Caveolae participation in signal transduction. The ever-increasing number of signaling molecules found concentrated in caveolar and similar domains has generated a re-evaluation of signaling mechanisms. It has become apparent that a large number of signaling molecules and, consequently, their associated signaling cascades are present in or can be recruited to caveolae. The current list includes receptor [e.g., EGF, platelet-derived growth factor (PDGF), insulin receptor] and nonreceptor (i.e., src family members) tyrosine kinases (32, 85, 144, 268). G protein-coupled receptors and their associated G protein effectors and targets (somatostatin receptor, adenosine receptor, heterotrimeric G proteins, adenylcyclase, PI-3 kinase) (4, 6, 185) and membrane transporters (aquaporin-1, Ca2+ ATPase, H+ ATPase, and IP3 receptor) (4, 103, 185) have been localized to caveolae. Recent evidence indicates caveolins can interact directly with and regulate the activity of a number of proteins, including endothelial nitric oxide synthetase, sonic hedgehog receptor, and src tyrosine kinase family members (182, 220, 230).
In addition to caveolin, caveolae are characterized by being highly enriched in cholesterol, sphingomyelin, and glycosphingolipids (70, 184; also reviewed in Ref. 27). As implied by the fluid-mosaic model, most of the lipid bilayer that forms the bulk of the biological membranes is in a fluid state at biological temperatures (229, 286). This is due to the presence of unsaturated (i.e., double bond-containing) fatty acyl groups at the sn-2 position of most cellular glycerophospholipids. Sphingomyelin and glycosphingolipids tend to possess longer, more saturated (i.e., no double bonds) fatty acyl groups, and bilayers composed of these compounds adopt an immobile gel (i.e., solid-like) phase at biological temperatures. Membrane cholesterol tends to dissolve into these regions, creating a liquid-ordered phase with fluidity intermediate between gel-ordered and mobile-fluid phase. Due to their lipid composition, caveolae can be isolated on the basis of buoyant density on suitable gradients (216, 232). The most commonly applied technique involves collecting those microdomains that are insoluble in 1% Triton X-100 at 0-4°C on 0-30% sucrose gradients (35, 163).
Overexpression of caveolin-1 in certain cells leads to an increase in the number of caveolae (68, 231). However, detergent-insoluble cholesterol-sphingomyelin-glycosphingolipid-enriched domains (DIGs) can readily be isolated not only from synthetic liposomes but also from cells such as those from the hematopoietic system and alveolar type II cells, which do not express detectable amounts of caveolins. Sequestering membrane cholesterol with saponin, filipin, or methylcyclodextrin leads to the loss of caveolar invaginations and the detergent-insoluble microdomains (4, 27, 146). It should be apparent that DIGs isolated from cells or tissues could arise either from caveolin-containing invaginations (or vesicles) or from caveolin-lacking cholesterol-sphingomyelin-rich microdomains. Further, both kinds of DIGs could arise from membranes other than the plasma membrane. The noncaveolin-containing microdomains would be flat and lack distinct morphological features by electron microscopy. Such caveolin-deficient cholesterol-sphingomyelin-glycosphingolipid-rich microdomains are often referred to as "rafts" [although readers should be cautious because some literature considers caveolae to be a particular type of raft, so that "rafts" refers to both caveolae (i.e., vesicular) and noncaveolin-containing (i.e., morphologically indistinct) cholesterol-sphingolipid-rich regions]. The concept of membrane rafts is still controversial because DIGs could be generated during detergent processing of the plasma and other membranes. However, considerable evidence has accumulated demonstrating that certain membrane proteins cluster in biological membranes, consistent with the presence of specialized regions (4, 26, 27).
The dynamics between noncaveolin-containing rafts and caveolin-dependent caveolae are still being clarified. Recent studies using model green fluorescent protein (GFP) and GFP-yes chimeric proteins (yes is a src-family member) indicate fatty-acid modifications such as NH2-terminal acylation contribute to protein sorting into cholesterol-sphingolipid-rich domains, but protein-protein interactions are also important (153, 154). Lipid rafts appear enriched in GPI-anchored proteins, including alkaline phosphatase, 5'-nucleotidase, CD14, folate receptor, and the prion protein (4, 94, 169). The heterotrimeric G proteins Gi and Gs concentrate in lipid rafts, whereas Gq appears to interact specifically with caveolin and is retained in caveolae (181). This agrees with other evidence showing that, although rafts and caveolae share a requirement for sphingolipids and cholesterol, they are biochemically and functionally as well as morphologically distinct.
Evidence for pulmonary rafts/caveolae.
DIGs isolated from whole rat lung contain a high proportion of total
cholesterol, sphingomyelin plus PC (measured as
phosphocholine-containing lipids), and caveolin-1
(176). Rat lung DIGs are
also enriched in GPI proteins, 5'-nucleotidase, and alkaline
phosphatase. Although PKC-, PKC-µ, and PKC-
II were
predominantly present in detergent-soluble fractions, a small
proportion of PKC-
II and PKC-µ colocalized with caveolin-1. Rat
lung DIGs account for ~0.25% of total tissue protein.
|
![]() |
CELLULAR FUNCTIONS FOR PAP1 AND LPP |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
PAP1. This review will close with comments on potential physiological roles for phosphatidate phosphohydrolase activities in the lung, with special emphasis on areas considered important for investigation at the present time. Lung contains two distinct kinds of phosphatidate phosphohydrolase. PAP1 is an Mg2+-dependent, cytosolic protein that translocates to the ER and possibly other membranes under the influence of free fatty acids and acyl CoAs. Whether other cellular agents affect PAP1 activity should be determined. PAP1 functions in glycerolipid synthesis by hydrolyzing newly synthesized PA to DAG (Fig. 1). PAP1 activity is essential in lung for the de novo synthesis of phospholipids for cellular membranes and for the production of PC and 1,2-dipalmitoyl-sn-3-phosphatidylcholine for pulmonary surfactant by type II cells (42). As PAP1 has not yet been purified or cloned, it has not been possible to ascertain whether more than one form of this enzyme is present in mammalian tissues, including lung. In addition, little information is available concerning the manner in which this enzyme is regulated.
PAP1 involvement in de novo glycerophospholipid and neutral lipid synthesis does not limit its potential participation in signal transduction. Chemoattractant stimulation of human polymorphonuclear leukocytes promotes PAP1 translocation from soluble to particulate fractions, indicating a potential role in signal transduction (251). Coimmunoprecipitation studies show PAP1 associates with the EGF receptor in unstimulated human epidermoid A431 cells but interacts with PKC-LPP. LPP is a six-transmembrane-domain, glycosylated, NEM-insensitive enzyme of ~28 kDa with its active site directed away from the cytosol. The bulk of the LPP activity in rat lung localizes with the plasma membrane (174). Lung and type II cells express the mRNAs for LPP, LPP1a, and LPP3 (175). Rat lung also expresses relatively high levels of LPP1b mRNA, which codes for a truncated protein of 30 amino acids. The physiological significance of LPP1b in lung is unknown but provides an interesting problem for future investigation.
In the rat, the specific activity of LPP in the lung ranks among the highest of any tissue. LPP activity is only slightly enriched in type II cells and is very low in lung fibroblasts (174). It would be of considerable interest to determine whether pulmonary type I cells, which exhibit abundant caveolae and express high levels of caveolin, possess high LPP activity, as would be predicted from studies on type II cell trans-/dedifferentiation in culture. In addition, it would be of interest to know whether type I cell caveolae possess LPP1/1a, as do MLE12 and MLE15 mouse type II cells, or LPP3, as do HEK 293 and Swiss 3T3 cells. Such information could provide a basis for investigating potential biochemical functions of type I cells. Part of the remaining "excess" LPP activity in whole lung tissue could arise from endothelial cells, which have high phosphatidate phosphohydrolase activity in other tissues (59). Whether pulmonary endothelial cells function in transcytosis and/or potocytosis as do endothelial cells in other organs is not presently known. In this regard, although direct evidence is lacking, it is conceivable that type I cells participate in transcytosis and/or potocytosis.Other lipid phosphate hydrolases. Future studies on LPP function must also consider the possibility of additional enzymes that hydrolyze lipid phosphates. Recently Mandala et al. (149) cloned the cDNA for a 43-kDa mammalian sphingosinephosphate phosphatase (mSPP1) from the mouse brain that degrades S1P, but not PA, LPA, or C1P. The cDNA showed little overall homology to LPP phosphohydrolases except for conserved amino acids at the predicted active site. Hydropathy plots suggest 8-10 potential membrane-spanning helices and a single potential glycosylation site, which was situated opposite the active site. This would place the active site on the cytosolic side of cellular membranes. Northern blot analysis found high mRNA expression in the liver and kidney, with relatively low levels in the lung and undetectable levels in skeletal muscle. Interestingly, the mouse tissue mRNA levels are discordant with S1P degradation activities previously assayed in the rat, where the brain, for example, possesses 3.5-fold greater activity than the liver or kidney (149). Overexpression of mSPP1 in NIH 3T3 fibroblasts leads to cell death, apparently by apoptosis.
An LPA phosphohydrolase whose activity was stimulated by the gonadotropin-releasing hormone antagonist buserelin has been reported in ovarian cells (101). This enzyme is Mg2+ independent and NEM insensitive but was not inhibited by S1P or C1P. A potentially distinct LPA-specific phosphohydrolase has been reported in the bovine brain (86) and PAM 212 keratinocytes (276). In addition, evidence has been obtained for an NEM-insensitive LPA phosphohydrolase, inhibited by S1P but not by PA, in nuclear membranes isolated from rabbit cerebral cortex neurons (7). Such an enzyme would be able to cleave LPA generated by a combination of nuclear PLA2 and PLD activities. (8, 11, 39). Further evidence for phosphohydrolases distinct from LPP arises from yeast, where separate proteins coded by diacylglycerol pyrophosphate (DGPP) phosphatase and LPP genes demonstrate the ability to degrade a number of lipid phosphates (83, 246). DGPP phosphatase is present in bacteria, and DGPP and DGPP phosphatase have critical signaling functions in plants (170). Although DGPP has not been identified in mammalian tissues, exogenous DGPP activates arachidonate secretion by mouse macrophages (9) and DGPP antagonizes LPA3 and, to a lesser extent, LPA1 receptor interactions (66, 245). It is recognized that DGPP is hydrolyzed by LPP (52, 107), but this does not preclude expression of a separate DGPP phosphatase activity as occurs in yeast. When considered together, these reports would suggest the likelihood of finding other members of the LPP family or other phosphohydrolases capable of degrading some lipid phosphates. Determining whether such proposed phosphohydrolases will be present and play distinct roles in the lung requires further investigation.LPPs and lung function. Gaseous exchange, the major physiological function of the lung, requires a sufficient alveolar surface area, appropriate surface tension as influenced by pulmonary surfactant, and a capillary bed positioned to deliver CO2 and accept O2. S1P is known to influence angiogenesis through action via S1P1 receptors on smooth muscle and endothelial cells (59, 145). Although the S1P1 receptor clearly plays a role in angiogenesis, the functions of the remaining S1P and LPA receptors are less defined, and the importance of these receptors in lung must still be examined in further detail.
Surface tensions compatible with normal lung function are dependent not only on adequate surfactant synthesis but additionally on alveolar surfactant homeostasis in terms of surfactant secretion and recycling back into type II cells (79, 137, 152). Surfactant secretion can be stimulated in isolated alveolar type II cells through three distinct signal transduction pathways: 1) A2
|
![]() |
FINAL CONSIDERATIONS |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
This review has attempted to summarize the present state of our understanding of LPP in the lung and respiratory tract. LPPs have been implicated in growth and differentiation of endothelial cells, fibroblasts, and type II cells. LPP localization in caveolae/rafts suggests participation in signaling cascades and in regulation of the effects of phospholipid growth factors. Although direct evidence is still inadequate, LPPs likely play important roles in lung vascularization and alveolar remodeling. LPPs have been implicated in surfactant secretion by type II cells and could have specific roles in type I cells.
Future studies should address LPP functions during development and in lung injury. Part of the difficulty experienced in this area can be attributed to the lack of cell culture systems that faithfully duplicate in vivo alterations. It should be apparent that application of gene-introducing and gene-disrupting recombinant technologies could provide novel insights. As indicated in LPP IN SIGNALING PLATFORMS, caveolin-1 gene ablation results in thickened alveolar septa (56, 201). Ablation of the LPP2 gene, which in the human is expressed primarily in the brain, pancreas, and placenta, was without obvious phenotype (282). However, ablation of the S1P1 receptor results in fetal demise at approximately embryonic day 12.5, most likely as a result of disturbed blood vessel development (145).
These initial applications of gene recombinant technology to the study of lipid phosphate functions demonstrate potential for yielding new insights into pathways mediating lung morphogenesis and repair. The availability of promoters such as the SP-C promoter for type II cells and the CC10 promoter for Clara cells has allowed introduction of unique cell-specific alterations in gene and cDNA expression. The doxycycline-regulatable reverse tetracycline transactivator system has proven particularly useful for regulating expression in epithelial cells of the conducting and peripheral airways. This tetracycline transactivator system has recently been utilized for introducing conditional expression of FGF family members into Clara and type II cells (270). Future studies employing these approaches in combination with the Cre-lox recombinase system (158, 159, 218) to regulate specific LPP isoform expression spatially and temporally hold considerable promise for investigating physiological functions of these lipid phosphohydrolases in lung.
![]() |
NOTE ADDED IN PROOF |
---|
The nomenclature for lysophospholipid receptors is described in the following review: Chun J, Goetzl EJ, Hla T, Igarashi Y, Lynch KR, Moolenaar W, Pyne S, and Tigyi G. International Union of Pharmacology. XXXIV. Lysophospholipid receptor nomenclature. Pharmacol Rev 54: 265-269, 2002.
![]() |
ACKNOWLEDGEMENTS |
---|
The authors thank Dr. Lin Zhao, Jon Faulkner, Anne Brickenden, Dr. Karina Rodriguez Capote, and Dr. David Brindley for reading the text and Dr. Jian Wang and Ross Ridsdale for discussions on the chromosome locations of human LPPs. We apologize to those whose work was not included for reasons of space.
![]() |
FOOTNOTES |
---|
These studies were supported by a Canadian Institutes of Health Research group grant.
Present address for M. Nanjundan: Scripps Research Institute - MEM-275, Dept. of Molecular and Experimental Medicine, 10550 No. Torrey Pines Rd., La Jolla, CA 92037.
Address for reprint requests and other correspondence: F. Possmayer, Dept. of Ob/Gyn, Univ. of Western Ontario, 339 Windermere Rd. London, Ontario Canada N6A 5A5 (E-mail: fpossmay{at}uwo.ca).
10.1152/ajplung.00029.2002
![]() |
REFERENCES |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
1.
Adamson, IY,
Hedgecock C,
and
Bowden DH.
Epithelial cell-fibroblast interactions in lung injury and repair.
Am J Pathol
137:
385-392,
1990[Abstract].
2.
Agranoff, BW.
Hydrolysis of long-chain alkyl phosphates and phosphatidic acid by an enzyme purified from pig brain.
J Lipid Res
3:
190-196,
1962
3.
An, S.
Molecular identification and characterization of G protein-coupled receptors for lysophosphatidic acid and sphingosine 1-phosphate.
Ann NY Acad Sci
905:
25-33,
2000
4.
Anderson, RG.
The caveolae membrane system.
Annu Rev Biochem
67:
199-225,
1998[ISI][Medline].
5.
Arbibe, L,
Koumanov K,
Vial D,
Rougeot C,
Faure G,
Havet N,
Longacre S,
Vargaftig BB,
Bereziat G,
Voelker DR,
Wolf C,
and
Touqui L.
Generation of lyso-phospholipids from surfactant in acute lung injury is mediated by type-II phospholipase A2 and inhibited by a direct surfactant protein A-phospholipase A2 protein interaction.
J Clin Invest
102:
1152-1160,
1998
6.
Arnal, JF,
Dinh-Xuan AT,
Pueyo M,
Darblade B,
and
Rami J.
Endothelium-derived nitric oxide and vascular physiology and pathology.
Cell Mol Life Sci
55:
1078-1087,
1999[ISI][Medline].
7.
Baker, RR,
and
Chang H.
A metabolic path for the degradation of lysophosphatidic acid, an inhibitor of lysophosphatidylcholine lysophospholipase, in neuronal nuclei of cerebral cortex.
Biochim Biophys Acta
1483:
58-68,
2000[ISI][Medline].
8.
Balboa, MA,
Balsinde J,
Dennis EA,
and
Insel PA.
A phospholipase D-mediated pathway for generating diacylglycerol in nuclei from Madin-Darby canine kidney cells.
J Biol Chem
270:
11738-11740,
1995
9.
Balboa, MA,
Balsinde J,
Dillon DA,
Carman GM,
and
Dennis EA.
Proinflammatory macrophage-activating properties of the novel phospholipid diacylglycerol pyrophosphate.
J Biol Chem
274:
522-526,
1999
10.
Balboa, MA,
and
Insel PA.
Nuclear phospholipase D in Madin-Darby canine kidney cells. Guanosine 5'-O-(thiotriphosphate)-stimulated activation is mediated by RhoA and is downstream of protein kinase C.
J Biol Chem
270:
29843-29847,
1995
11.
Baldassare, JJ,
Jarpe MB,
Alferes L,
and
Raben DM.
Nuclear translocation of RhoA mediates the mitogen-induced activation of phospholipase D involved in nuclear envelope signal transduction.
J Biol Chem
272:
4911-4914,
1997
12.
Barila, D,
Murgia C,
Nobili F,
Gaetani S,
and
Perozzi G.
Subtractive hybridization cloning of novel genes differentially expressed during intestinal development.
Eur J Biochem
223:
701-709,
1994[Abstract].
13.
Berridge, MJ.
Cell signalling. A tale of two messengers.
Nature
365:
388-389,
1993[ISI][Medline].
14.
Berridge, MJ.
Inositol trisphosphate and calcium signalling.
Nature
361:
315-325,
1993[ISI][Medline].
15.
Berridge, MJ.
Inositol trisphosphate and diacylglycerol: two interacting second messengers.
Annu Rev Biochem
56:
159-193,
1987[ISI][Medline].
16.
Berridge, MJ,
and
Irvine RF.
Inositol phosphates and cell signalling.
Nature
341:
197-205,
1989[ISI][Medline].
17.
Bevers, EM,
Comfurius P,
Dekkers DW,
and
Zwaal RF.
Lipid translocation across the plasma membrane of mammalian cells.
Biochim Biophys Acta
1439:
317-330,
1999[ISI][Medline].
18.
Bocckino, SB,
and
Exton JH.
Phosphatidic acid.
In: Lipid Second Messengers, edited by Bell RM,
Exton JH,
and Prescott SM.. New York: Plenum, 1996, p. 75-123.
19.
Bocckino, SB,
Wilson PB,
and
Exton JH.
Phosphatidate-dependent protein phosphorylation.
Proc Natl Acad Sci USA
88:
6210-6213,
1991[Abstract].
20.
Bowton, DL,
Seeds MC,
Fasano MB,
Goldsmith B,
and
Bass DA.
Phospholipase A2 and arachidonate increase in bronchoalveolar lavage fluid after inhaled antigen challenge in asthmatics.
Am J Respir Crit Care Med
155:
421-425,
1997[Abstract].
21.
Brazil, DP,
and
Hemmings BA.
Ten years of protein kinase B signalling: a hard Akt to follow.
Trends Biochem Sci
26:
657-664,
2001[ISI][Medline].
22.
Brindley, DN.
Intracellular translocation of phosphatidate phosphohydrolase and its possible role in the control of glycerolipid synthesis.
Prog Lipid Res
23:
115-133,
1984[ISI][Medline].
23.
Brindley, DN.
Phosphatidate phosphohydrolase: its role in glycerolipid synthesis.
In: Phosphatidate Phosphohydrolase, edited by Brindley DN.. Boca Raton, FL: CRC, 1988, p. 21-77.
24.
Brindley, DN,
English D,
Pilquil C,
Buri K,
and
Ling Z-C.
Lipid phosphate phosphatases regulate signal transduction through glycerolipids and sphingolipids.
Biochim Biophys Acta
1582:
33-44,
2002[ISI][Medline].
25.
Brindley, DN,
and
Waggoner DW.
Mammalian lipid phosphate phosphohydrolases.
J Biol Chem
273:
24281-24284,
1998
26.
Brown, DA,
and
London E.
Functions of lipid rafts in biological membranes.
Annu Rev Cell Dev Biol
14:
111-136,
1998[ISI][Medline].
27.
Brown, DA,
and
London E.
Structure of detergent-resistant membrane domains: does phase separation occur in biological membranes?
Biochem Biophys Res Commun
240:
1-7,
1997[ISI][Medline].
28.
Bui, KC,
Buckley S,
Wu F,
Uhal B,
Joshi I,
Liu J,
Hussain M,
Makhoul I,
and
Warburton D.
Induction of A- and D-type cyclins and cdc2 kinase activity during recovery from short-term hyperoxic lung injury.
Am J Physiol Lung Cell Mol Physiol
268:
L625-L635,
1995
29.
Bui, KC,
Wu F,
Buckley S,
Wu L,
Williams R,
Carbonaro-Hall D,
Hall FL,
and
Warburton D.
Cyclin A expression in normal and transformed alveolar epithelial cells.
Am J Respir Cell Mol Biol
9:
115-125,
1993[ISI][Medline].
30.
Campbell, L,
Hollins AJ,
Al-Eid A,
Newman GR,
von Ruhland C,
and
Gumbleton M.
Caveolin-1 expression and caveolae biogenesis during cell transdifferentiation in lung alveolar epithelial primary cultures.
Biochem Biophys Res Commun
262:
744-751,
1999[ISI][Medline].
31.
Carman, GM.
Phosphatidate phosphatases and diacylglycerol pyrophosphate phosphatases in Saccharomyces cerevisiae and Escherichia coli.
Biochim Biophys Acta
1348:
45-55,
1997[ISI][Medline].
32.
Carpenter, G.
The EGF receptor: a nexus for trafficking and signaling.
Bioessays
22:
697-707,
2000[ISI][Medline].
33.
Casola, PG,
and
Possmayer F.
Pulmonary phosphatidic acid phosphatase. Properties of membrane-bound phosphatidate-dependent phosphatidic acid phosphatase in rat lung.
Biochim Biophys Acta
574:
212-225,
1979[ISI][Medline].
34.
Chan, C,
and
Goldkorn T.
Ceramide path in human lung cell death.
Am J Respir Cell Mol Biol
22:
460-468,
2000
35.
Chang, WJ,
Ying YS,
Rothberg KG,
Hooper NM,
Turner AJ,
Gambliel HA,
De Gunzburg J,
Mumby SM,
Gilman AG,
and
Anderson RG.
Purification and characterization of smooth muscle cell caveolae.
J Cell Biol
126:
127-138,
1994[Abstract].
36.
Chilton, FH,
Averill FJ,
Hubbard WC,
Fonteh AN,
Triggiani M,
and
Liu MC.
Antigen-induced generation of lyso-phospholipids in human airways.
J Exp Med
183:
2235-2245,
1996[Abstract].
37.
Chuang, TH,
Bohl BP,
and
Bokoch GM.
Biologically active lipids are regulators of Rac. GDI complexation.
J Biol Chem
268:
26206-26211,
1993
38.
Chun, J,
Weiner JA,
Fukushima N,
Contos JJ,
Zhang G,
Kimura Y,
Dubin A,
Ishii I,
Hecht JH,
Akita C,
and
Kaushal D.
Neurobiology of receptor-mediated lysophospholipid signaling. From the first lysophospholipid receptor to roles in nervous system function and development.
Ann NY Acad Sci
905:
110-117,
2000
39.
Clark, JM,
Hodgkin MN,
and
Wakelam MJ.
HL60 nuclei lacking the nuclear double membrane contain a PLD activity which is insensitive to the ADP-ribosylation factor.
Biochem Soc Trans
25:
S590,
1997[ISI][Medline].
40.
Cockcroft, S.
Signalling roles of mammalian phospholipase D1 and D2.
Cell Mol Life Sci
58:
1674-1687,
2001[ISI][Medline].
41.
Coleman, R,
and
Hubscher G.
Metabolism of phospholipids. V. Studies on phosphatidic acid phosphatase.
Biochim Biophys Acta
56:
479-490,
1962.
42.
Coleman, RA,
Lewin TM,
and
Muoio DM.
Physiological and nutritional regulation of enzymes of triacylglycerol synthesis.
Annu Rev Nutr
20:
77-103,
2000[ISI][Medline].
43.
Colley, WC,
Sung TC,
Roll R,
Jenco J,
Hammond SM,
Altshuller Y,
Bar-Sagi D,
Morris AJ,
and
Frohman MA.
Phospholipase D2, a distinct phospholipase D isoform with novel regulatory properties that provokes cytoskeletal reorganization.
Curr Biol
7:
191-201,
1997[ISI][Medline].
44.
Czarny, M,
Fiucci G,
Lavie Y,
Banno Y,
Nozawa Y,
and
Liscovitch M.
Phospholipase D2: functional interaction with caveolin in low-density membrane microdomains.
FEBS Lett
467:
326-332,
2000[ISI][Medline].
45.
Czarny, M,
Lavie Y,
Fiucci G,
and
Liscovitch M.
Localization of phospholipase D in detergent-insoluble, caveolin-rich membrane domains. Modulation by caveolin-1 expression and caveolin-182-101.
J Biol Chem
274:
2717-2724,
1999
46.
Daniel, LW,
Sciorra VA,
and
Ghosh S.
Phospholipase D, tumor promoters, proliferation and prostaglandins.
Biochim Biophys Acta
1439:
265-276,
1999[ISI][Medline].
47.
Danielli, JF,
and
Davson H.
A contribution to the theory of permeability of thin films.
J Cell Comp Physiol
5:
495-508,
1935.
48.
De Paepe, ME,
Rubin LP,
Jude C,
Lesieur-Brooks AM,
Mills DR,
and
Luks FI.
Fas ligand expression coincides with alveolar cell apoptosis in late-gestation fetal lung development.
Am J Physiol Lung Cell Mol Physiol
279:
L967-L976,
2000
49.
DeMello, DE,
Mahmoud S,
Padfield PJ,
and
Hoffmann JW.
Generation of an immortal differentiated lung type-II epithelial cell line from the adult H-2K(b)tsA58 transgenic mouse.
In Vitro Cell Dev Biol Anim
36:
374-382,
2000[ISI][Medline].
50.
Dempsey, EC,
Newton AC,
Mochly-Rosen D,
Fields AP,
Reyland ME,
Insel PA,
and
Messing RO.
Protein kinase C isozymes and the regulation of diverse cell responses.
Am J Physiol Lung Cell Mol Physiol
279:
L429-L438,
2000
51.
Dennis, EA.
Phospholipase A2 in eicosanoid generation.
Am J Respir Crit Care Med
161:
S32-S35,
2000
52.
Dillon, DA,
Chen X,
Zeimetz GM,
Wu WI,
Waggoner DW,
Dewald J,
Brindley DN,
and
Carman GM.
Mammalian Mg2+-independent phosphatidate phosphatase (PAP2) displays diacylglycerol pyrophosphate phosphatase activity.
J Biol Chem
272:
10361-10366,
1997
53.
Divecha, N,
and
Irvine RF.
Phospholipid signaling.
Cell
80:
269-278,
1995[ISI][Medline].
54.
Dobbs, LG.
Isolation and culture of alveolar type II cells.
Am J Physiol Lung Cell Mol Physiol
258:
L134-L147,
1990
55.
Dobbs, LG,
Pian MS,
Maglio M,
Dumars S,
and
Allen L.
Maintenance of the differentiated type II cell phenotype by culture with an apical air surface.
Am J Physiol Lung Cell Mol Physiol
273:
L347-L354,
1997
56.
Drab, M,
Verkade P,
Elger M,
Kasper M,
Lohn M,
Lauterbach B,
Menne J,
Lindschau C,
Mende F,
Luft FC,
Schedl A,
Haller H,
and
Kurzchalia TV.
Loss of caveolae, vascular dysfunction, and pulmonary defects in caveolin-1 gene-disrupted mice.
Science
293:
2449-2452,
2001
57.
Egawa, K,
Yoshiwara M,
Shibanuma M,
and
Nose K.
Isolation of a novel ras-recision gene that is induced by hydrogen peroxide from a mouse osteoblastic cell line, MC3T3-E1.
FEBS Lett
372:
74-77,
1995[ISI][Medline].
58.
Engelman, JA,
Zhang XL,
Razani B,
Pestell RG,
and
Lisanti MP.
p42/44 MAP kinase-dependent and -independent signaling pathways regulate caveolin-1 gene expression. Activation of Ras-MAP kinase and protein kinase a signaling cascades transcriptionally down-regulates caveolin-1 promoter activity.
J Biol Chem
274:
32333-32341,
1999
59.
English, D,
Kovala AT,
Welch Z,
Harvey KA,
Siddiqui RA,
Brindley DN,
and
Garcia JG.
Induction of endothelial cell chemotaxis by sphingosine 1-phosphate and stabilization of endothelial monolayer barrier function by lysophosphatidic acid, potential mediators of hematopoietic angiogenesis.
J Hematother Stem Cell Res
8:
627-634,
1999[ISI][Medline].
60.
Exton, JH.
New developments in phospholipase D.
J Biol Chem
272:
15579-15582,
1997
61.
Exton, JH.
Regulation of phospholipase D.
Biochim Biophys Acta
1439:
121-133,
1999[ISI][Medline].
62.
Exton, JH.
Signaling through phosphatidylcholine breakdown.
J Biol Chem
265:
1-4,
1990
63.
Ferguson, MA,
Brimacombe JS,
Brown JR,
Crossman A,
Dix A,
Field RA,
Guther ML,
Milne KG,
Sharma DK,
and
Smith TK.
The GPI biosynthetic pathway as a therapeutic target for African sleeping sickness.
Biochim Biophys Acta
1455:
327-340,
1999[ISI][Medline].
64.
Fielding, CJ,
and
Fielding PE.
Cholesterol and caveolae: structural and functional relationships.
Biochim Biophys Acta
1529:
210-222,
2000[ISI][Medline].
65.
Fine, A,
Janssen-Heininger Y,
Soultanakis RP,
Swisher SG,
and
Uhal BD.
Apoptosis in lung pathophysiology.
Am J Physiol Lung Cell Mol Physiol
279:
L423-L427,
2000
66.
Fischer, DJ,
Nusser N,
Virag T,
Yokoyama K,
Wang D,
Baker DL,
Bautista D,
Parrill AL,
and
Tigyi G.
Short-chain phosphatidates are subtype-selective antagonists of lysophosphatidic acid receptors.
Mol Pharmacol
60:
776-784,
2001
67.
Fortunati, E,
Bout A,
Zanta MA,
Valerio D,
and
Scarpa M.
In vitro and in vivo gene transfer to pulmonary cells mediated by cationic liposomes.
Biochim Biophys Acta
1306:
55-62,
1996[ISI][Medline].
68.
Fra, AM,
Williamson E,
Simons K,
and
Parton RG.
De novo formation of caveolae in lymphocytes by expression of VIP21-caveolin.
Proc Natl Acad Sci USA
92:
8655-8659,
1995[Abstract].
69.
Frohman, MA,
Sung TC,
and
Morris AJ.
Mammalian phospholipase D structure and regulation.
Biochim Biophys Acta
1439:
175-186,
1999[ISI][Medline].
70.
Ge, M,
Field KA,
Aneja R,
Holowka D,
Baird B,
and
Freed JH.
Electron spin resonance characterization of liquid ordered phase of detergent-resistant membranes from RBL-2H3 cells.
Biophys J
77:
925-933,
1999
71.
Ghosh, S,
and
Bell RM.
Regulation of Raf-1 kinase by interaction with the lipid second messenger, phosphatidic acid.
Biochem Soc Trans
25:
561-565,
1997[ISI][Medline].
72.
Gijon, MA,
and
Leslie CC.
Regulation of arachidonic acid release and cytosolic phospholipase A2 activation.
J Leukoc Biol
65:
330-336,
1999[Abstract].
73.
Gijon, MA,
Spencer DM,
and
Leslie CC.
Recent advances in the regulation of cytosolic phospholipase A(2).
Adv Enzyme Regul
40:
255-268,
2000[ISI][Medline].
74.
Gobran, LI,
Xu ZX,
Lu Z,
and
Rooney SA.
P2u purinoceptor stimulation of surfactant secretion coupled to phosphatidylcholine hydrolysis in type II cells.
Am J Physiol Lung Cell Mol Physiol
267:
L625-L633,
1994
75.
Gobran, LI,
Xu ZX,
and
Rooney SA.
PKC isoforms and other signaling proteins involved in surfactant secretion in developing rat type II cells.
Am J Physiol Lung Cell Mol Physiol
274:
L901-L907,
1998
76.
Goetzl, EJ,
and
An S.
Diversity of cellular receptors and functions for the lysophospholipid growth factors lysophosphatidic acid and sphingosine 1-phosphate.
FASEB J
12:
1589-1598,
1998
77.
Gorter, E,
and
Grendel F.
On bimolecular layers of lipoids on the chromocytes of blood.
J Exp Med
41:
439-443,
1925.
78.
Griese, M,
Gobran LI,
and
Rooney SA.
ATP-stimulated inositol phospholipid metabolism and surfactant secretion in rat type II pneumocytes.
Am J Physiol Lung Cell Mol Physiol
260:
L586-L593,
1991
79.
Gross, NJ.
Extracellular metabolism of pulmonary surfactant: the role of a new serine protease.
Annu Rev Physiol
57:
135-150,
1995[ISI][Medline].
80.
Ha, KS,
Yeo EJ,
and
Exton JH.
Lysophosphatidic acid activation of phosphatidylcholine-hydrolysing phospholipase D and actin polymerization by a pertussis toxin-sensitive mechanism.
Biochem J
303:
55-59,
1994[ISI][Medline].
81.
Hammond, SM,
Altshuller YM,
Sung TC,
Rudge SA,
Rose K,
Engebrecht J,
Morris AJ,
and
Frohman MA.
Human ADP-ribosylation factor-activated phosphatidylcholine-specific phospholipase D defines a new and highly conserved gene family.
J Biol Chem
270:
29640-29643,
1995
82.
Hammond, SM,
Jenco JM,
Nakashima S,
Cadwallader K,
Gu Q,
Cook S,
Nozawa Y,
Prestwich GD,
Frohman MA,
and
Morris AJ.
Characterization of two alternately spliced forms of phospholipase D1. Activation of the purified enzymes by phosphatidylinositol 4,5-bisphosphate, ADP-ribosylation factor, and Rho family monomeric GTP-binding proteins and protein kinase C-alpha.
J Biol Chem
272:
3860-3868,
1997
83.
Han, GS,
Johnston CN,
Chen X,
Athenstaedt K,
Daum G,
and
Carman GM.
Regulation of the Saccharomyces cerevisiae DPP1-encoded diacylglycerol pyrophosphate phosphatase by zinc.
J Biol Chem
276:
10126-10133,
2001
84.
Hanahan, DJ,
and
Chaikoff IL.
A new phospholipid-splitting enzyme specific for the ester linkage between the nitrogenous base and the phosphoric acid grouping.
J Biol Chem
169:
699-705,
1947
85.
Harada, S,
Smith RM,
and
Jarett L.
Mechanisms of nuclear translocation of insulin.
Cell Biochem Biophys
31:
307-319,
1999[ISI][Medline].
86.
Hiroyama, M,
and
Takenawa T.
Purification and characterization of a lysophosphatidic acid-specific phosphatase.
Biochem J
336:
483-489,
1998[ISI][Medline].
87.
Hite, RD,
Seeds MC,
Jacinto RB,
Balasubramanian R,
Waite M,
and
Bass D.
Hydrolysis of surfactant-associated phosphatidylcholine by mammalian secretory phospholipases A2.
Am J Physiol Lung Cell Mol Physiol
275:
L740-L747,
1998
88.
Hla, T,
Lee MJ,
Ancellin N,
Liu CH,
Thangada S,
Thompson BD,
and
Kluk M.
Sphingosine-1-phosphate: extracellular mediator or intracellular second messenger?
Biochem Pharmacol
58:
201-207,
1999[ISI][Medline].
89.
Hla, T,
Lee MJ,
Ancellin N,
Thangada S,
Liu CH,
Kluk M,
Chae SS,
and
Wu MT.
Sphingosine-1-phosphate signaling via the EDG-1 family of G-protein-coupled receptors.
Ann NY Acad Sci
905:
16-24,
2000
90.
Hla, T,
and
Maciag T.
An abundant transcript induced in differentiating human endothelial cells encodes a polypeptide with structural similarities to G-protein-coupled receptors.
J Biol Chem
265:
9308-9313,
1990
91.
Hodgkin, MN,
Pettitt TR,
Martin A,
Michell RH,
Pemberton AJ,
and
Wakelam MJO
Diacylglycerols and phosphatidates: which molecular species are intracellular messengers?
Trends Biochem Sci
23:
200-204,
1998[ISI][Medline].
92.
Hokin, LE,
Hokin MR,
and
Mathison D.
Phosphatidic acid phosphatase in erythrocyte membrane.
Biochim Biophys Acta
67:
485-497,
1963[ISI].
93.
Hooks, SB,
Ragan SP,
and
Lynch KR.
Identification of a novel human phosphatidic acid phosphatase type 2 isoform.
FEBS Lett
427:
188-192,
1998[ISI][Medline].
94.
Hooper, NM.
Detergent-insoluble glycosphingolipid/cholesterol-rich membrane domains, lipid rafts and caveolae.
Mol Membr Biol
16:
145-156,
1999[ISI][Medline].
95.
Hope, MJ,
Mui B,
Ansell S,
and
Ahkong QF.
Cationic lipids, phosphatidylethanolamine and the intracellular delivery of polymeric, nucleic acid-based drugs.
Mol Membr Biol
15:
1-14,
1998[ISI][Medline].
96.
Houle, MG,
and
Bourgoin S.
Regulation of phospholipase D by phosphorylation-dependent mechanisms.
Biochim Biophys Acta
1439:
135-149,
1999[ISI][Medline].
97.
Hubscher, G,
Brindley DN,
Smith ME,
and
Sedgwick B.
Stimulation of biosynthesis of glyceride.
Nature
216:
449-453,
1967[ISI][Medline].
98.
Hulit, J,
Bash T,
Fu M,
Galbiati F,
Albanese C,
Sage DR,
Schlegel A,
Zhurinsky J,
Shtutman M,
Ben-Ze'ev A,
Lisanti MP,
and
Pestell RG.
The cyclin D1 gene is transcriptionally repressed by caveolin-1.
J Biol Chem
275:
21203-21209,
2000
99.
Hussain, MM.
A proposed model for the assembly of chylomicrons.
Atherosclerosis
148:
1-15,
2000[ISI][Medline].
100.
Igarashi, J,
and
Michel T.
Agonist-modulated targeting of the EDG-1 receptor to plasmalemmal caveolae. eNOS activation by sphingosine 1-phosphate and the role of caveolin-1 in sphingolipid signal transduction.
J Biol Chem
275:
32363-32370,
2000
101.
Imai, A,
Furui T,
Tamaya T,
and
Mills GB.
A gonadotropin-releasing hormone-responsive phosphatase hydrolyses lysophosphatidic acid within the plasma membrane of ovarian cancer cells.
J Clin Endocrinol Metab
85:
3370-3375,
2000
102.
Irvine, RF,
and
Schell MJ.
Back in the water: the return of the inositol phosphates.
Nat Rev Mol Cell Biol
2:
327-338,
2001[ISI][Medline].
103.
Isshiki, M,
and
Anderson RG.
Calcium signal transduction from caveolae.
Cell Calcium
26:
201-208,
1999[ISI][Medline].
104.
Jackowski, S,
and
Rock CO.
Stimulation of phosphatidylinositol 4,5-bisphosphate phospholipase C activity by phosphatidic acid.
Arch Biochem Biophys
268:
516-524,
1989[ISI][Medline].
105.
Jamal, Z,
Martin A,
Gomez-Munoz A,
Hales P,
Chang E,
Russell JC,
and
Brindley DN.
Phosphatidate phosphohydrolases in liver, heart and adipose tissue of the JCR:LA corpulent rat and the lean genotypes: implications for glycerolipid synthesis and signal transduction.
Int J Obes Relat Metab Disord
16:
789-799,
1992[Medline].
106.
Jamdar, SC,
and
Fallon HJ.
Glycerolipid synthesis in rat adipose tissue. II. Properties and distribution of phosphatidate phosphatase.
J Lipid Res
14:
517-524,
1973
107.
Jasinska, R,
Zhang QX,
Pilquil C,
Singh I,
Xu J,
Dewald J,
Dillon DA,
Berthiaume LG,
Carman GM,
Waggoner DW,
and
Brindley DN.
Lipid phosphate phosphohydrolase-1 degrades exogenous glycerolipid and sphingolipid phosphate esters.
Biochem J
340:
677-686,
1999[ISI][Medline].
108.
Jenkins, GH,
Fisette PL,
and
Anderson RA.
Type I phosphatidylinositol 4-phosphate 5-kinase isoforms are specifically stimulated by phosphatidic acid.
J Biol Chem
269:
11547-11554,
1994
109.
Jiang, Y,
Lu Z,
Zang Q,
and
Foster DA.
Regulation of phosphatidic acid phosphohydrolase by epidermal growth factor. Reduced association with the EGF receptor followed by increased association with protein kinase Cepsilon.
J Biol Chem
271:
29529-29532,
1996
110.
Johnson, CA,
Balboa MA,
Balsinde J,
and
Dennis EA.
Regulation of cyclooxygenase-2 expression by phosphatidate phosphohydrolase in human amnionic WISH cells.
J Biol Chem
274:
27689-27693,
1999
111.
Johnston, JM,
and
Bearden JH.
Intestinal phosphatidate phosphatase.
Biochim Biophys Acta
56:
365-367,
1962.
112.
Johnston, JM,
Rao GA,
Lowe PA,
and
Schwarz BE.
The nature of the stimulatory role of the supernatant fraction on triglyceride synthesis by the alpha-glycerophosphate pathway.
Lipids
2:
14-20,
1967[ISI].
113.
Jones, D,
Morgan C,
and
Cockcroft S.
Phospholipase D and membrane traffic. Potential roles in regulated exocytosis, membrane delivery and vesicle budding.
Biochim Biophys Acta
1439:
229-244,
1999[ISI][Medline].
114.
Jones, GA,
and
Carpenter G.
The regulation of phospholipase C-gamma 1 by phosphatidic acid. Assessment of kinetic parameters.
J Biol Chem
268:
20845-20850,
1993
115.
Jones, LG,
Ella KM,
Bradshaw CD,
Gause KC,
Dey M,
Wisehart-Johnson AE,
Spivey EC,
and
Meier KE.
Activations of mitogen-activated protein kinases and phospholipase D in A7r5 vascular smooth muscle cells.
J Biol Chem
269:
23790-23799,
1994
116.
Kai, M,
Sakane F,
Imai S,
Wada I,
and
Kanoh H.
Molecular cloning of a diacylglycerol kinase isozyme predominantly expressed in human retina with a truncated and inactive enzyme expression in most other human cells.
J Biol Chem
269:
18492-18498,
1994
117.
Kai, M,
Wada I,
Imai S,
Sakane F,
and
Kanoh H.
Cloning and characterization of two human isozymes of Mg++-independent phosphatidic acid phosphatase.
J Biol Chem
272:
24572-24578,
1997
118.
Kai, M,
Wada I,
Imai S,
Sakane F,
and
Kanoh H.
Identification and cDNA cloning of 35-kDa phosphatidic acid phosphatase (type 2) bound to plasma membranes. Polymerase chain reaction amplification of mouse H2O2-inducible hic53 clone yielded the cDNA encoding phosphatidic acid phosphatase.
J Biol Chem
271:
18931-18938,
1996
119.
Kanfer, JN.
The base exchange enzymes and phospholipase D of mammalian tissue.
Can J Biochem
58:
1370-1380,
1980[ISI][Medline].
120.
Kanoh, H,
Kai M,
and
Wada I.
Molecular characterization of the type 2 phosphatidic acid phosphatase.
Chem Phys Lipids
98:
119-126,
1999[ISI][Medline].
121.
Kasper, M,
Reimann T,
Hempel U,
Wenzel KW,
Bierhaus A,
Schuh D,
Dimmer V,
Haroske G,
and
Muller M.
Loss of caveolin expression in type I pneumocytes as an indicator of subcellular alterations during lung fibrogenesis.
Histochem Cell Biol
109:
41-48,
1998[ISI][Medline].
122.
Katayama, K,
Kodaki T,
Nagamachi Y,
and
Yamashita S.
Cloning, differential regulation and tissue distribution of alternatively spliced isoforms of ADP-ribosylation-factor-dependent phospholipase D from rat liver.
Biochem J
329:
647-652,
1998[ISI][Medline].
123.
Kates, M.
Hydrolysis of lecithin by plant plastid enzymes.
Can J Biochem Physiol
33:
575-589,
1955[ISI].
124.
Kennedy, EP.
Biosynthesis of complex lipids.
Fed Proc
20:
934-940,
1961[ISI].
125.
Kennedy, EP.
The biosynthesis of phospholipids.
In: Lipids and Membranes: Past, Present, and Future, edited by Op den Kamp JAF,
Roelofsen B,
and Wirtz KWA. Amsterdam: Elsevier Science, 1986, p. 171-206.
126.
Khan, WA,
Blobe GC,
Richards AL,
and
Hannun YA.
Identification, partial purification, and characterization of a novel phospholipid-dependent and fatty acid-activated protein kinase from human platelets.
J Biol Chem
269:
9729-9735,
1994
127.
Kim, JH,
Han JM,
Lee S,
Kim Y,
Lee TG,
Park JB,
Lee SD,
Suh PG,
and
Ryu SH.
Phospholipase D1 in caveolae: regulation by protein kinase Calpha and caveolin-1.
Biochemistry
38:
3763-3769,
1999[ISI][Medline].
128.
Kim, Y,
Han JM,
Han BR,
Lee KA,
Kim JH,
Lee BD,
Jang IH,
Suh PG,
and
Ryu SH.
Phospholipase D1 is phosphorylated and activated by protein kinase C in caveolin-enriched microdomains within the plasma membrane.
J Biol Chem
275:
13621-13627,
2000
129.
Kitagawa, M,
Mukai H,
Shibata H,
and
Ono Y.
Purification and characterization of a fatty acid-activated protein kinase (PKN) from rat testis.
Biochem J
310:
657-664,
1995[ISI][Medline].
130.
Kodaki, T,
and
Yamashita S.
Cloning, expression, and characterization of a novel phospholipase D complementary DNA from rat brain.
J Biol Chem
272:
11408-11413,
1997
131.
Krieger, M.
Charting the fate of the "good cholesterol": identification and characterization of the high-density lipoprotein receptor SR-BI.
Annu Rev Biochem
68:
523-558,
1999[ISI][Medline].
132.
Lee, SD,
Lee BD,
Han JM,
Kim JH,
Kim Y,
Suh PG,
and
Ryu SH.
Phospholipase D2 activity suppresses hydrogen peroxide-induced apoptosis in PC12 cells.
J Neurochem
75:
1053-1059,
2000[ISI][Medline].
133.
Leff, AR.
Discovery of leukotrienes and development of antileukotriene agents.
Ann Allergy Asthma Immunol
86:
4-8,
2001[ISI][Medline].
134.
Leslie, CC,
McCormick-Shannon K,
Shannon JM,
Garrick B,
Damm D,
Abraham JA,
and
Mason RJ.
Heparin-binding EGF-like growth factor is a mitogen for rat alveolar type II cells.
Am J Respir Cell Mol Biol
16:
379-387,
1997[Abstract].
135.
Leung, DW,
Tompkins CK,
and
White T.
Molecular cloning of two alternatively spliced forms of human phosphatidic acid phosphatase cDNAs that are differentially expressed in normal and tumor cells.
DNA Cell Biol
17:
377-385,
1998[ISI][Medline].
136.
Levine, L,
and
Hassid A.
Effects of phorbol-12,13-diesters on prostaglandin production and phospholipase activity in canine kidney (MDCK) cells.
Biochem Biophys Res Commun
79:
477-484,
1977[ISI][Medline].
137.
Lewis, JF,
and
Veldhuizen RA.
Factors influencing efficacy of exogenous surfactant in acute lung injury.
Biol Neonate
67:
48-60,
1995[ISI][Medline].
138.
Limatola, C,
Schaap D,
Moolenaar WH,
and
van Blitterswijk WJ.
Phosphatidic acid activation of protein kinase C-zeta overexpressed in COS cells: comparison with other protein kinase C isotypes and other acidic lipids.
Biochem J
304:
1001-1008,
1994[ISI][Medline].
139.
Liscovitch, M,
Czarny M,
Fiucci G,
Lavie Y,
and
Tang X.
Localization and possible functions of phospholipase D isozymes.
Biochim Biophys Acta
1439:
245-263,
1999[ISI][Medline].
140.
Liscovitch, M,
Czarny M,
Fiucci G,
and
Tang X.
Phospholipase D: molecular and cell biology of a novel gene family.
Biochem J
345:
401-415,
2000[ISI][Medline].
141.
Liscovitch, M,
and
Lavie Y.
Multidrug resistance: a role for cholesterol efflux pathways?
Trends Biochem Sci
25:
530-534,
2000[ISI][Medline].
142.
Liu, H,
Sugiura M,
Nava VE,
Edsall LC,
Kono K,
Poulton S,
Milstien S,
Kohama T,
and
Spiegel S.
Molecular cloning and functional characterization of a novel mammalian sphingosine kinase type 2 isoform.
J Biol Chem
275:
19513-19520,
2000
143.
Liu, J,
Oh P,
Horner T,
Rogers RA,
and
Schnitzer JE.
Organized endothelial cell surface signal transduction in caveolae distinct from glycosylphosphatidylinositol-anchored protein microdomains.
J Biol Chem
272:
7211-7222,
1997
144.
Liu, P,
Ying Y,
Ko YG,
and
Anderson RG.
Localization of platelet-derived growth factor-stimulated phosphorylation cascade to caveolae.
J Biol Chem
271:
10299-10303,
1996
145.
Liu, Y,
Wada R,
Yamashita T,
Mi Y,
Deng CX,
Hobson JP,
Rosenfeldt HM,
Nava VE,
Chae SS,
Lee MJ,
Liu CH,
Hla T,
Spiegel S,
and
Proia RL.
Edg-1, the G protein-coupled receptor for sphingosine-1-phosphate, is essential for vascular maturation.
J Clin Invest
106:
951-961,
2000
146.
London, E,
and
Brown DA.
Insolubility of lipids in triton X-100: physical origin and relationship to sphingolipid/cholesterol membrane domains (rafts).
Biochim Biophys Acta
1508:
182-195,
2000[ISI][Medline].
147.
Mallampalli, RK,
Peterson EJ,
Carter AB,
Salome RG,
Mathur SN,
and
Koretzky GA.
TNF- increases ceramide without inducing apoptosis in alveolar type II epithelial cells.
Am J Physiol Lung Cell Mol Physiol
276:
L481-L490,
1999
148.
Mandala, SM.
Sphingosine-1-phosphate phosphatases.
Prostaglandins
64:
143-156,
2001[ISI][Medline].
149.
Mandala, SM,
Thornton R,
Galve-Roperh I,
Poulton S,
Peterson C,
Olivera A,
Bergstrom J,
Kurtz MB,
and
Spiegel S.
Molecular cloning and characterization of a lipid phosphohydrolase that degrades sphingosine-1-phosphate and induces cell death.
Proc Natl Acad Sci USA
97:
7859-7864,
2000
150.
Marcoz, P,
Nemoz G,
Prigent AF,
and
Lagarde M.
Phosphatidic acid stimulates the rolipram-sensitive cyclic nucleotide phosphodiesterase from rat thymocytes.
Biochim Biophys Acta
1176:
129-136,
1993[ISI][Medline].
151.
Martin, A,
Gomez-Munoz A,
Waggoner DW,
Stone JC,
and
Brindley DN.
Decreased activities of phosphatidate phosphohydrolase and phospholipase D in ras and tyrosine kinase (fps) transformed fibroblasts.
J Biol Chem
268:
23924-23932,
1993
152.
Mason, RJ,
and
Voelker DR.
Regulatory mechanisms of surfactant secretion.
Biochim Biophys Acta
1408:
226-240,
1998[ISI][Medline].
153.
McCabe, JB,
and
Berthiaume LG.
Functional roles for fatty acylated amino-terminal domains in subcellular localization.
Mol Biol Cell
10:
3771-3786,
1999
154.
McCabe, JB,
and
Berthiaume LG.
N-terminal protein acylation confers localization to cholesterol, sphingolipid-enriched membranes but not to lipid rafts/caveolae.
Mol Biol Cell
12:
3601-3617,
2001
155.
McPhail, LC,
Qualliotine-Mann D,
and
Waite KA.
Cell-free activation of neutrophil NADPH oxidase by a phosphatidic acid-regulated protein kinase.
Proc Natl Acad Sci USA
92:
7931-7935,
1995[Abstract].
156.
McPhail, LC,
Waite KA,
Regier DS,
Nixon JB,
Qualliotine-Mann D,
Zhang WX,
Wallin R,
and
Sergeant S.
A novel protein kinase target for the lipid second messenger phosphatidic acid.
Biochim Biophys Acta
1439:
277-290,
1999[ISI][Medline].
157.
Mendelson, CR.
Role of transcription factors in fetal lung development and surfactant protein gene expression.
Annu Rev Physiol
62:
875-915,
2000[ISI][Medline].
158.
Metzger, D,
and
Chambon P.
Site- and time-specific gene targeting in the mouse.
Methods
24:
71-80,
2001[Medline].
159.
Meuwissen, R,
Linn SC,
van der Valk M,
Mooi WJ,
and
Berns A.
Mouse model for lung tumorigenesis through Cre/lox controlled sporadic activation of the K-Ras oncogene.
Oncogene
20:
6551-6558,
2001[ISI][Medline].
160.
Mineo, C,
Anderson RG,
and
White MA.
Physical association with ras enhances activation of membrane-bound raf (RafCAAX).
J Biol Chem
272:
10345-10348,
1997
161.
Mitchell, MP,
Brindley DN,
and
Hubscher G.
Properties of phosphatidate phosphohydrolase.
Eur J Biochem
18:
214-220,
1971[ISI][Medline].
162.
Miyata, Y,
and
Yahara I.
p53-independent association between SV40 large T antigen and the major cytosolic heat shock protein, HSP90.
Oncogene
19:
1477-1484,
2000[ISI][Medline].
163.
Moldovan, NI,
Heltianu C,
Simionescu N,
and
Simionescu M.
Ultrastructural evidence of differential solubility in Triton X-100 of endothelial vesicles and plasma membrane.
Exp Cell Res
219:
309-313,
1995[ISI][Medline].
164.
Moolenaar, WH.
Bioactive lysophospholipids and their G protein-coupled receptors.
Exp Cell Res
253:
230-238,
1999[ISI][Medline].
165.
Moolenaar, WH.
Development of our current understanding of bioactive lysophospholipids.
Ann NY Acad Sci
905:
1-10,
2000
166.
Moritz, A,
De Graan PN,
Gispen WH,
and
Wirtz KW.
Phosphatidic acid is a specific activator of phosphatidylinositol-4-phosphate kinase.
J Biol Chem
267:
7207-7210,
1992
167.
Morrice, NA,
Gabrielli B,
Kemp BE,
and
Wettenhall RE.
A cardiolipin-activated protein kinase from rat liver structurally distinct from the protein kinases C.
J Biol Chem
269:
20040-20046,
1994
168.
Muller, G.
The molecular mechanism of the insulin-mimetic/sensitizing activity of the antidiabetic sulfonylurea drug Amaryl.
Mol Med
6:
907-933,
2000[ISI][Medline].
169.
Muller, G,
and
Frick W.
Signalling via caveolin: involvement in the cross-talk between phosphoinositolglycans and insulin.
Cell Mol Life Sci
56:
945-970,
1999[ISI][Medline].
170.
Munnik, T.
Phosphatidic acid: an emerging plant lipid second messenger.
Trends Plant Sci
6:
227-233,
2001[ISI][Medline].
171.
Murakami, M,
Nakatani Y,
Kuwata H,
and
Kudo I.
Cellular components that functionally interact with signaling phospholipase A(2)s.
Biochim Biophys Acta
1488:
159-166,
2000[ISI][Medline].
172.
Murayama, T,
and
Ui M.
Phosphatidic acid may stimulate membrane receptors mediating adenylate cyclase inhibition and phospholipid breakdown in 3T3 fibroblasts.
J Biol Chem
262:
5522-5529,
1987
173.
Nakanishi, H,
and
Exton JH.
Purification and characterization of the zeta isoform of protein kinase C from bovine kidney.
J Biol Chem
267:
16347-16354,
1992
174.
Nanjundan, M,
and
Possmayer F.
Characterization of the pulmonary N-ethylmaleimide-insensitive phosphatidate phosphohydrolase.
Exp Lung Res
26:
361-381,
2000[ISI][Medline].
175.
Nanjundan, M,
and
Possmayer F.
Molecular cloning and expression of pulmonary lipid phosphate phosphohydrolases.
Am J Physiol Lung Cell Mol Physiol
281:
L1484-L1493,
2001
176.
Nanjundan, M,
and
Possmayer F.
Pulmonary lipid phosphate phosphohydrolase in plasma membrane signalling platforms.
Biochem J
358:
637-646,
2001[ISI][Medline].
177.
Nava, VE,
Lacana E,
Poulton S,
Liu H,
Sugiura M,
Kono K,
Milstien S,
Kohama T,
and
Spiegel S.
Functional characterization of human sphingosine kinase-1.
FEBS Lett
473:
81-84,
2000[ISI][Medline].
178.
Newman, GR,
Campbell L,
von Ruhland C,
Jasani B,
and
Gumbleton M.
Caveolin and its cellular and subcellular immunolocalisation in lung alveolar epithelium: implications for alveolar epithelial type I cell function.
Cell Tissue Res
295:
111-120,
1999[ISI][Medline].
179.
Oh, P,
McIntosh DP,
and
Schnitzer JE.
Dynamin at the neck of caveolae mediates their budding to form transport vesicles by GTP-driven fission from the plasma membrane of endothelium.
J Cell Biol
141:
101-114,
1998
180.
Oh, P,
and
Schnitzer JE.
Immunoisolation of caveolae with high affinity antibody binding to the oligomeric caveolin cage. Toward understanding the basis of purification.
J Biol Chem
274:
23144-23154,
1999
181.
Oh, P,
and
Schnitzer JE.
Segregation of heterotrimeric G proteins in cell surface microdomains. G(q) binds caveolin to concentrate in caveolae, whereas G(i) and G(s) target lipid rafts by default.
Mol Biol Cell
12:
685-698,
2001
182.
Oh, P,
and
Schnitzer JE.
Segregation of heterotrimeric G proteins in cell surface microdomains. G(q) binds caveolin to concentrate in caveolae, whereas G(i) and G(s) target lipid rafts by default.
Mol Biol Cell
12:
685-698,
2001
183.
Okamura, S,
and
Yamashita S.
Purification and characterization of phosphatidylcholine phospholipase D from pig lung.
J Biol Chem
269:
31207-31213,
1994
184.
Ostermeyer, AG,
Beckrich BT,
Ivarson KA,
Grove KE,
and
Brown DA.
Glycosphingolipids are not essential for formation of detergent-resistant membrane rafts in melanoma cells. Methyl-beta-cyclodextrin does not affect cell surface transport of a GPI-anchored protein.
J Biol Chem
274:
34459-34466,
1999
185.
Ostrom, RS,
Post SR,
and
Insel PA.
Stoichiometry and compartmentation in G protein-coupled receptor signaling: implications for therapeutic interventions involving G(s).
J Pharmacol Exp Ther
294:
407-412,
2000
186.
Overton, E.
Über die allgemeinen osmotischen Eigenschaften der Zelle, ihre vermütliche Ursachen und ihre Bedeutung für die Physiologie.
Vierteljahrsschrift der Naturforschende Gesellschaft Zürich
44:
88-114,
1899.
187.
Park, SK,
Provost JJ,
Bae CD,
Ho WT,
and
Exton JH.
Cloning and characterization of phospholipase D from rat brain.
J Biol Chem
272:
29263-29271,
1997
188.
Park, WY,
Park JS,
Cho KA,
Kim DI,
Ko YG,
Seo JS,
and
Park SC.
Up-regulation of caveolin attenuates epidermal growth factor signaling in senescent cells.
J Biol Chem
275:
20847-20852,
2000
189.
Parmentier, J-H,
Muthalif MM,
Saeed AE,
and
Malik KU.
Phospholipase D activation by norepinephrine is mediated by 12(S)-, 15(S)-, and 20-hydroxyeicosatetraenoic acids generated by stimulation of cytosolic phospholipase A2.
J Biol Chem
276:
15704-15711,
2001
190.
Pessin, MS,
and
Raben DM.
Molecular species analysis of 1,2-diglycerides stimulated by alpha-thrombin in cultured fibroblasts.
J Biol Chem
264:
8729-8738,
1989
191.
Pettitt, TR,
McDermott M,
Saqib KM,
Shimwell N,
and
Wakelam MJ.
Phospholipase D1b and D2a generate structurally identical phosphatidic acid species in mammalian cells.
Biochem J
360:
707-715,
2001[ISI][Medline].
192.
Pettitt, TR,
and
Wakelam MJ.
Bombesin stimulates distinct time-dependent changes in the sn-1,2-diradylglycerol molecular species profile from Swiss 3T3 fibroblasts as analysed by 3,5-dinitrobenzoyl derivatization and h.p.l.c. separation.
Biochem J
289:
487-495,
1993[ISI][Medline].
193.
Pettitt, TR,
and
Wakelam MJ.
Diacylglycerol kinase epsilon, but not zeta, selectively removes polyunsaturated diacylglycerol, inducing altered protein kinase C distribution in vivo.
J Biol Chem
274:
36181-36186,
1999
194.
Possmayer, F.
Pulmonary phosphatidate phosphohydrolase and its relation to the surfactant system of the lung.
In: Phosphatidate Phosphohydrolase, edited by Brindley DN.. Boca Raton, FL: CRC, 1988, p. 39-118.
195.
Prescott, SM,
Zimmerman GA,
Stafforini DM,
and
McIntyre TM.
Platelet-activating factor and related lipid mediators.
Annu Rev Biochem
69:
419-445,
2000[ISI][Medline].
196.
Pyne, S,
and
Pyne NJ.
Sphingosine 1-phosphate signalling in mammalian cells.
Biochem J
349:
385-402,
2000[ISI][Medline].
197.
Qualliotine-Mann, D,
Agwu DE,
Ellenburg MD,
McCall CE,
and
McPhail LC.
Phosphatidic acid and diacylglycerol synergize in a cell-free system for activation of NADPH oxidase from human neutrophils.
J Biol Chem
268:
23843-23849,
1993
198.
Racine, C,
Belanger M,
Hirabayashi H,
Boucher M,
Chakir J,
and
Couet J.
Reduction of caveolin 1 gene expression in lung carcinoma cell lines.
Biochem Biophys Res Commun
255:
580-586,
1999[ISI][Medline].
199.
Racke, K,
Hammermann R,
and
Juergens UR.
Potential role of EDG receptors and lysophospholipids as their endogenous ligands in the respiratory tract.
Pulm Pharmacol Ther
13:
99-114,
2000[ISI][Medline].
200.
Raggers, RJ,
Pomorski T,
Holthuis JC,
Kalin N,
and
van Meer G.
Lipid traffic: the ABC of transbilayer movement.
Traffic
1:
226-234,
2000[ISI][Medline].
201.
Razani, B,
Engelman JA,
Wang XB,
Schubert W,
Zhang XL,
Marks CB,
Macaluso F,
Russell RG,
Li M,
Pestell RG,
Di Vizio D,
Hou H, Jr,
Kneitz B,
Lagaud G,
Christ GJ,
Edelmann W,
and
Lisanti MP.
Caveolin-1 null mice are viable but show evidence of hyperproliferative and vascular abnormalities.
J Biol Chem
276:
38121-38138,
2001
202.
Rebecchi, MJ,
and
Pentyala SN.
Structure, function, and control of phosphoinositide-specific phospholipase C.
Physiol Rev
80:
1291-1335,
2000
203.
Renkonen, O.
Mono- and dimethyl phosphatidates from different subtypes of choline and ethanolamine glycerophosphatides.
Biochim Biophys Acta
152:
114-135,
1968[ISI][Medline].
204.
Rice, WR,
Burton FM,
and
Fiedeldey DT.
Cloning and expression of the alveolar type II cell P2u-purinergic receptor.
Am J Respir Cell Mol Biol
12:
27-32,
1995[Abstract].
205.
Ritter, TE,
Fajardo O,
Matsue H,
Anderson RG,
and
Lacey SW.
Folate receptors targeted to clathrin-coated pits cannot regulate vitamin uptake.
Proc Natl Acad Sci USA
92:
3824-3828,
1995
206.
Roberts, R,
Sciorra VA,
and
Morris AJ.
Human type 2 phosphatidic acid phosphohydrolases: substrate specificity of the type 2a, 2b, and 2c enzymes and cell surface activity of the 2a isoform.
J Biol Chem
273:
22059-22067,
1998
207.
Robertson, JD.
A review of membrane structure with perspectives on certain transmembrane channels.
Adv Neurol
31:
419-477,
1981[Medline].
208.
Ron, D,
and
Kazanietz MG.
New insights into the regulation of protein kinase C and novel phorbol ester receptors.
FASEB J
13:
1658-1676,
1999
209.
Rooney, SA.
Regulation of surfactant secretion.
Comp Biochem Physiol A Mol Integr Physiol
129:
233-243,
2001[ISI][Medline].
210.
Rooney, SA.
Regulation of surfactant-associated phospholipid synthesis and secretion.
In: Fetal and Neonatal Physiology, edited by Polin RA,
and Fox WW.. Philadelphia, PA: WB Saunders, 1997, p. 1283-1299.
211.
Rooney, SA,
and
Gobran LI.
Activation of phospholipase D in rat type II pneumocytes by ATP and other surfactant secretagogues.
Am J Physiol Lung Cell Mol Physiol
264:
L133-L140,
1993
212.
Rooney, SA,
Young SL,
and
Mendelson CR.
Molecular and cellular processing of lung surfactant.
FASEB J
8:
957-967,
1994
213.
Roth, MG,
Bi K,
Ktistakis NT,
and
Yu S.
Phospholipase D as an effector for ADP-ribosylation factor in the regulation of vesicular traffic.
Chem Phys Lipids
98:
141-152,
1999[ISI][Medline].
214.
Saito, M,
and
Kanfer J.
Phosphatidohydrolase activity in a solubilized preparation from rat brain particulate fraction.
Arch Biochem Biophys
169:
318-323,
1975[ISI][Medline].
215.
Sane, AC,
Mendenhall T,
and
Bass DA.
Secretory phospholipase A2 activity is elevated in bronchoalveolar lavage fluid after ovalbumin sensitization of guinea pigs.
J Leukoc Biol
60:
704-709,
1996[Abstract].
216.
Sargiacomo, M,
Sudol M,
Tang Z,
and
Lisanti MP.
Signal transducing molecules and glycosyl-phosphatidylinositol-linked proteins form a caveolin-rich insoluble complex in MDCK cells.
J Cell Biol
122:
789-807,
1993[Abstract].
217.
Sarkar, S,
Miwa N,
Kominami H,
Igarashi N,
Hayashi S,
Okada T,
Jahangeer S,
and
Nakamura S.
Regulation of mammalian phospholipase D2: interaction with and stimulation by GM2 activator.
Biochem J
359:
599-604,
2001[ISI][Medline].
218.
Sauer, B.
Inducible gene targeting in mice using the Cre/lox system.
Methods
14:
381-392,
1998[ISI][Medline].
219.
Savany, A,
Abriat C,
Nemoz G,
Lagarde M,
and
Prigent AF.
Activation of a cyclic nucleotide phosphodiesterase 4 (PDE4) from rat thymocytes by phosphatidic acid.
Cell Signal
8:
511-516,
1996[ISI][Medline].
220.
Schlegel, A,
and
Lisanti MP.
Caveolae and their coat proteins, the caveolins: from electron microscopic novelty to biological launching pad.
J Cell Physiol
186:
329-337,
2001[ISI][Medline].
221.
Schmidt, A,
Wolde M,
Thiele C,
Fest W,
Kratzin H,
Podtelejnikov AV,
Witke W,
Huttner WB,
and
Soling HD.
Endophilin I mediates synaptic vesicle formation by transfer of arachidonate to lysophosphatidic acid.
Nature
401:
133-141,
1999[ISI][Medline].
222.
Schmitz, G,
Kaminski WE,
and
Orso E.
ABC transporters in cellular lipid trafficking.
Curr Opin Lipidol
11:
493-501,
2000[ISI][Medline].
223.
Schnitzer, JE,
Liu J,
and
Oh P.
Endothelial caveolae have the molecular transport machinery for vesicle budding, docking, and fusion including VAMP, NSF, SNAP, annexins, and GTPases.
J Biol Chem
270:
14399-14404,
1995
224.
Sciorra, VA,
and
Morris AJ.
Sequential actions of phospholipase D and phosphatidic acid phosphohydrolase 2b generate diglyceride in mammalian cells.
Mol Biol Cell
10:
3863-3876,
1999
225.
Sedgwick, B,
and
Hubscher G.
Metabolism of phospholipids. IX. Phosphatidate phosphohydrolase in rat liver.
Biochim Biophys Acta
106:
63-77,
1965[ISI][Medline].
226.
Shannon, JM,
Jennings SD,
and
Nielsen LD.
Modulation of alveolar type II cell differentiated function in vitro.
Am J Physiol Lung Cell Mol Physiol
262:
L427-L436,
1992
227.
Shin, JS,
and
Abraham SN.
Co-option of endocytic functions of cellular caveolae by pathogens.
Immunology
102:
2-7,
2001[ISI][Medline].
228.
Siddhanta, A,
and
Shields D.
Secretory vesicle budding from the trans-Golgi network is mediated by phosphatidic acid levels.
J Biol Chem
273:
17995-17998,
1998
229.
Singer, SJ,
and
Nicolson GL.
The fluid mosaic model of the structure of cell membranes.
Science
175:
720-731,
1972[ISI][Medline].
230.
Smart, EJ,
Graf GA,
McNiven MA,
Sessa WC,
Engelman JA,
Scherer PE,
Okamoto T,
and
Lisanti MP.
Caveolins, liquid-ordered domains, and signal transduction.
Mol Cell Biol
19:
7289-7304,
1999
231.
Smart, EJ,
Ying Y,
Donzell WC,
and
Anderson RG.
A role for caveolin in transport of cholesterol from endoplasmic reticulum to plasma membrane.
J Biol Chem
271:
29427-35,
1996
232.
Smart, EJ,
Ying YS,
and
Anderson RG.
Hormonal regulation of caveolae internalization.
J Cell Biol
131:
929-938,
1995[Abstract].
233.
Smith, ME,
and
Hubscher G.
The biosynthesis of glycerides by mitochondria from rat liver. The requirement for a soluble protein.
Biochem J
101:
308-316,
1966[ISI][Medline].
234.
Smith, ME,
Sedgwick B,
Brindley DN,
and
Hubscher G.
The role of phosphatidate phosphohydrolase in glyceride biosynthesis.
Eur J Biochem
3:
70-77,
1967[ISI][Medline].
235.
Smith, SW,
Weiss SB,
and
Kennedy EP.
The enzymatic dephosphorylation of phosphatidic acids.
J Biol Chem
228:
915-922,
1957
236.
Snitko, Y,
Yoon ET,
and
Cho W.
High specificity of human secretory class II phospholipase A2 for phosphatidic acid.
Biochem J
321:
737-741,
1997[ISI][Medline].
237.
Spiegel, S,
and
Merrill AH, Jr.
Sphingolipid metabolism and cell growth regulation.
FASEB J
10:
1388-1397,
1996
238.
Steed, PM,
Clark KL,
Boyar WC,
and
Lasala DJ.
Characterization of human PLD2 and the analysis of PLD isoform splice variants.
FASEB J
12:
1309-1317,
1998
239.
Stein, Y,
Tietz A,
and
Shapiro B.
Glyceride synthesis by rat liver mitochondria.
Biochim Biophys Acta
26:
286-293,
1957[ISI].
240.
Stoeckenius, W,
and
Engelman DM.
Current models for the structure of biological membranes.
J Cell Biol
42:
613-646,
1969
241.
Stukey, J,
and
Carman GM.
Identification of a novel phosphatase sequence motif.
Protein Sci
6:
469-472,
1997
242.
Sturton, RG,
and
Brindley DN.
Factors controlling the metabolism of phosphatidate by phosphohydrolase and phospholipase A-type activities. Effects of magnesium, calcium and amphiphilic cationic drugs.
Biochim Biophys Acta
619:
494-505,
1980[ISI][Medline].
243.
Taki, T,
and
Kanfer JN.
Partial purification and properties of a rat brain phospholipase.
J Biol Chem
254:
9761-9765,
1979[Abstract].
244.
Tietz, DF,
and
Shapiro B.
The synthesis of glycerides in liver homogenates.
Biochim Biophys Acta
19:
374-375,
1956[ISI].
245.
Tigyi, G.
Selective ligands for lysophosphatidic acid receptor subtypes: gaining control over the endothelial differentiation gene family.
Mol Pharmacol
60:
1161-1164,
2001
246.
Toke, DA,
Bennett WL,
Dillon DA,
Wu WI,
Chen X,
Ostrander DB,
Oshiro J,
Cremesti A,
Voelker DR,
Fischl AS,
and
Carman GM.
Isolation and characterization of the Saccharomyces cerevisiae DPP1 gene encoding diacylglycerol pyrophosphate phosphatase.
J Biol Chem
273:
3278-3284,
1998
247.
Tokumura, A,
Nishioka Y,
Yoshimoto O,
Shinomiya J,
and
Fukuzawa K.
Substrate specificity of lysophospholipase D which produces bioactive lysophosphatidic acids in rat plasma.
Biochim Biophys Acta
1437:
235-245,
1999[ISI][Medline].
248.
Tolias, KF,
Couvillon AD,
Cantley LC,
and
Carpenter CL.
Characterization of a Rac1- and RhoGDI-associated lipid kinase signaling complex.
Mol Cell Biol
18:
762-770,
1998
249.
Tomic, S,
Greiser U,
Lammers R,
Kharitonenkov A,
Imyanitov E,
Ullrich A,
and
Bohmer FD.
Association of SH2 domain protein tyrosine phosphatases with the epidermal growth factor receptor in human tumor cells. Phosphatidic acid activates receptor dephosphorylation by PTP1C.
J Biol Chem
270:
21277-21284,
1995
250.
Topham, MK,
and
Prescott SM.
Mammalian diacylglycerol kinases, a family of lipid kinases with signaling functions.
J Biol Chem
274:
11447-11450,
1999
251.
Truett, AP, III,
Bocckino SB,
and
Murray JJ.
Regulation of phosphatidic acid phosphohydrolase activity during stimulation of human polymorphonuclear leukocytes.
FASEB J
6:
2720-2725,
1992
252.
Tsai, MH,
Roudebush M,
Dobrowolski S,
Yu CL,
Gibbs JB,
and
Stacey DW.
Ras GTPase-activating protein physically associates with mitogenically active phospholipids.
Mol Cell Biol
11:
2785-2793,
1991[ISI][Medline].
253.
Ueyama, T,
Ren Y,
Ohmori S,
Sakai K,
Tamaki N,
and
Saito N.
cDNA cloning of an alternative splicing variant of protein kinase C delta (PKC deltaIII), a new truncated form of PKCdelta, in rats.
Biochem Biophys Res Commun
269:
557-563,
2000[ISI][Medline].
254.
Valet, P,
Pages C,
Jeanneton O,
Daviaud D,
Barbe P,
Record M,
Saulnier-Blache JS,
and
Lafontan M.
Alpha2-adrenergic receptor-mediated release of lysophosphatidic acid by adipocytes. A paracrine signal for preadipocyte growth.
J Clin Invest
101:
1431-1438,
1998
255.
Van Blitterswijk, WJ,
and
Houssa B.
Properties and functions of diacylglycerol kinases.
Cell Signal
12:
595-605,
2000[ISI][Medline].
256.
Van Corven, EJ,
Groenink A,
Jalink K,
Eichholtz T,
and
Moolenaar WH.
Lysophosphatidate-induced cell proliferation: identification and dissection of signaling pathways mediated by G proteins.
Cell
59:
45-54,
1989[ISI][Medline].
257.
Van der Bend, RL,
de Widt J,
van Corven EJ,
Moolenaar WH,
and
van Blitterswijk WJ.
The biologically active phospholipid, lysophosphatidic acid, induces phosphatidylcholine breakdown in fibroblasts via activation of phospholipase D. Comparison with the response to endothelin.
Biochem J
285:
235-240,
1992[ISI][Medline].
258.
Waggoner, DW,
Gomez-Munoz A,
Dewald J,
and
Brindley DN.
Phosphatidate phosphohydrolase catalyzes the hydrolysis of ceramide 1-phosphate, lysophosphatidate, and sphingosine 1-phosphate.
J Biol Chem
271:
16506-16509,
1996
259.
Waggoner, DW,
Martin A,
Dewald J,
Gomez-Munoz A,
and
Brindley DN.
Purification and characterization of novel plasma membrane phosphatidate phosphohydrolase from rat liver.
J Biol Chem
270:
19422-19429,
1995
260.
Waggoner, DW,
Xu J,
Singh I,
Jasinska R,
Zhang QX,
and
Brindley DN.
Structural organization of mammalian lipid phosphate phosphatases: implications for signal transduction.
Biochim Biophys Acta
1439:
299-316,
1999[ISI][Medline].
261.
Wakelam, MJ.
Diacylglycerol-when is it an intracellular messenger?
Biochim Biophys Acta
1436:
117-126,
1998[ISI][Medline].
262.
Waksman, M,
Eli Y,
Liscovitch M,
and
Gerst JE.
Identification and characterization of a gene encoding phospholipase D activity in yeast.
J Biol Chem
271:
2361-2364,
1996
263.
Walton, PA,
and
Possmayer F.
The effects of Triton X-100 and chlorpromazine on the Mg2+-dependent and Mg2+-independent phosphatidate phosphohydrolase activities of rat lung.
Biochem J
261:
673-678,
1989[ISI][Medline].
264.
Walton, PA,
and
Possmayer F.
Mg2-dependent phosphatidate phosphohydrolase of rat lung: development of an assay employing a defined chemical substrate which reflects the phosphohydrolase activity measured using membrane-bound substrate.
Anal Biochem
151:
479-486,
1985[ISI][Medline].
265.
Walton, PA,
and
Possmayer F.
The role of Mg2+-dependent phosphatidate phosphohydrolase in pulmonary glycerolipid biosynthesis.
Biochim Biophys Acta
796:
364-372,
1984[ISI][Medline].
266.
Wang, X,
Xu L,
and
Zheng L.
Cloning and expression of phosphatidylcholine-hydrolyzing phospholipase D from Ricinus communis L.
J Biol Chem
269:
20312-20317,
1994
267.
Watanabe, G,
Saito Y,
Madaule P,
Ishizaki T,
Fujisawa K,
Morii N,
Mukai H,
Ono Y,
Kakizuka A,
and
Narumiya S.
Protein kinase N (PKN) and PKN-related protein rhophilin as targets of small GTPase Rho.
Science
271:
645-648,
1996[Abstract].
268.
Waterman, H,
and
Yarden Y.
Molecular mechanisms underlying endocytosis and sorting of ErbB receptor tyrosine kinases.
FEBS Lett
490:
142-152,
2001[ISI][Medline].
269.
Weigert, R,
Silletta MG,
Spano S,
Turacchio G,
Cericola C,
Colanzi A,
Senatore S,
Mancini R,
Polishchuk EV,
Salmona M,
Facchiano F,
Burger KN,
Mironov A,
Luini A,
and
Corda D.
CtBP/BARS induces fission of Golgi membranes by acylating lysophosphatidic acid.
Nature
402:
429-433,
1999[ISI][Medline].
270.
Whitsett, JA,
Glasser SW,
Tichelaar JW,
Perl AK,
Clark JC,
and
Wert SE.
Transgenic models for study of lung morphogenesis and repair: Parker B. Francis lecture.
Chest
120:
27S-30S,
2001[ISI][Medline].
271.
Wikenheiser, KA,
Clark JC,
Linnoila RI,
Stahlman MT,
and
Whitsett JA.
Simian virus 40 large T antigen directed by transcriptional elements of the human surfactant protein C gene produces pulmonary adenocarcinomas in transgenic mice.
Cancer Res
52:
5342-5352,
1992[Abstract].
272.
Wikenheiser, KA,
Vorbroker DK,
Rice WR,
Clark JC,
Bachurski CJ,
Oie HK,
and
Whitsett JA.
Production of immortalized distal respiratory epithelial cell lines from surfactant protein C/simian virus 40 large tumor antigen transgenic mice.
Proc Natl Acad Sci USA
90:
11029-11033,
1993[Abstract].
273.
Wilgram, GF,
and
Kennedy EP.
Intracellular distribution of some enzymes catalyzing reactions in the biosynthesis of complex lipids.
J Biol Chem
238:
2615-2619,
1963
274.
Winstead, MV,
Balsinde J,
and
Dennis EA.
Calcium-independent phospholipase A(2): structure and function.
Biochim Biophys Acta
1488:
28-39,
2000[ISI][Medline].
275.
Wu, F,
Buckley S,
Bui KC,
and
Warburton D.
Differential expression of cyclin D2 and cdc2 genes in proliferating and nonproliferating alveolar epithelial cells.
Am J Respir Cell Mol Biol
12:
95-103,
1995[Abstract].
276.
Xie, M,
and
Low MG.
Identification and characterization of an ecto-(lyso)phosphatidic acid phosphatase in PAM212 keratinocytes.
Arch Biochem Biophys
312:
254-259,
1994[ISI][Medline].
277.
Xie, Z,
Ho WT,
and
Exton JH.
Requirements and effects of palmitoylation of rat PLD1.
J Biol Chem
276:
9383-9391,
2001
278.
Xie, Z,
Ho WT,
and
Exton JH.
Conserved amino acids at the C-terminus of rat phospholipase D1 are essential for enzymatic activity.
Eur J Biochem
267:
7138-7146,
2000
279.
Yang, SF,
Freer S,
and
Benson AA.
Transphosphatidylation by phospholipase D.
J Biol Chem
242:
477-484,
1967
280.
Yokozeki, T,
Homma K,
Kuroda S,
Kikkawa U,
Ohno S,
Takahashi M,
Imahori K,
and
Kanaho Y.
Phosphatidic acid-dependent phosphorylation of a 29-kDa protein by protein kinase Calpha in bovine brain cytosol.
J Neurochem
71:
410-417,
1998[ISI][Medline].
281.
Yu, W,
Liu J,
Morrice NA,
and
Wettenhall RE.
Isolation and characterization of a structural homologue of human PRK2 from rat liver. Distinguishing substrate and lipid activator specificities.
J Biol Chem
272:
10030-10034,
1997
282.
Zhang, N,
Sundberg JP,
and
Gridley T.
Mice mutant for Ppap2c, a homolog of the germ cell migration regulator wunen, are viable and fertile.
Genesis
27:
137-140,
2000[ISI][Medline].
283.
Zhang, N,
Zhang J,
Purcell KJ,
Cheng Y,
and
Howard K.
The Drosophila protein Wunen repels migrating germ cells.
Nature
385:
64-67,
1997[ISI][Medline].
284.
Zhang, QX,
Pilquil CS,
Dewald J,
Berthiaume LG,
and
Brindley DN.
Identification of structurally important domains of lipid phosphate phosphatase-1: implications for its sites of action.
Biochem J
345:
181-184,
2000[ISI][Medline].
285.
Zhao, Z,
Shen SH,
and
Fischer EH.
Stimulation by phospholipids of a protein-tyrosine-phosphatase containing two src homology 2 domains.
Proc Natl Acad Sci USA
90:
4251-4255,
1993[Abstract].
286.
Zwaal, RFA,
Demel RA,
Roelofsen B,
and
van Deenen LLM
The lipid bilayer concept of cell membranes.
Trends Biochem Sci
1:
112-114,
1976[ISI].