1 Departments of Physiology and Biophysics and of Cell Biology, University of Alabama at Birmingham, Birmingham, Alabama, 35294-0005; and 2 Division of Nephrology and Hypertension, Department of Internal Medicine, University of Cincinnati Medical Center, Cincinnati, Ohio, 45267-0585
![]() |
ABSTRACT |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
During the past two decades, several cell membrane receptors, which preferentially bind extracellular nucleotides, and their analogs have been identified. These receptors, collectively known as nucleotide receptors or "purinergic" receptors, have been characterized and classified on the basis of their biological actions, their pharmacology, their molecular biology, and their tissue and cell distribution. For these receptors to have biological and physiological relevance, nucleotides must be released from cells. The field of extracellular ATP release and signaling is exploding, as assays to detect this biological process increase in number and ingenuity. Studies of ATP release have revealed a myriad of roles in local regulatory (autocrine or paracrine) processes in almost every tissue in the body. The regulatory mechanisms that these receptors control or modulate have physiological and pathophysiological roles and potential therapeutic applications. Only recently, however, have ATP release and nucleotide receptors been identified along the renal epithelium of the nephron. This work has set the stage for the study of their physiological and pathophysiological roles in the kidney. This review provides a comprehensive presentation of these issues, with a focus on the renal epithelium.
purinergic; adeonsine 5'-triphosphate; kidney; receptors; epithelia
![]() |
HISTORIC ASPECTS OF PURINERGIC SIGNALING |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
SEVENTY YEARS AGO, Drury and Szent-Gyorgyi (41) were the first to recognize the potent extracellular actions of purine nucleotides and nucleosides in mammalian heart. After this initial report, research was limited to the actions of adenosine and ATP on the cardiovascular system (37, 70, 123). In the early 1960s, a component of the autonomic nervous system that was neither adrenergic nor cholinergic was identified in several tissues. Moreover, aspects of the cardiovascular system also had an unknown neurotransmitter that was neither norepinephrine nor acetylcholine (21, 24). In the early 1970s, Burnstock (21, 24) proposed that the principal neurotransmitter that is released from these specialized nerves was ATP. On the basis of the accepted criteria for neurotransmitters, Burnstock (21, 24) proposed the term "purinergic" for these nonadrenergic and noncholinergic nerves. After Burnstock presented his proposal, the concept of extracellular purinergic neurotransmission was strengthened by a larger body of experimental evidence demonstrating the role of ATP as a neurotransmitter or cotransmitter with norepinephrine, acetylcholine, or other chemical mediators. In addition, knowledge about specific extracellular receptors that mediate the physiological effects of purine nucleotides and nucleosides began to emerge.
In 1980, Burnstock proposed the classification of purinergic receptors into two major groups, P1 and P2, depending on the preferential affinity to adenosine or ATP, respectively (22, 23). This classification also took into account the selective activation of adenylate cyclase by adenosine and the induction of prostaglandin synthesis by ATP. Subsequently, Lodos et al. (86) and Van Calker et al. (147) identified two subclasses of adenosine (P1) receptors. These subclasses were later termed "A1" and "A2" in the nomenclature. By using molecular homology cloning methods (157) and receptor binding assays (33), A3 and A4 subtypes of P1 receptors were also identified. The A4 subtype is still not validated fully. These adenosine receptor subtypes couple to different heterotrimeric G proteins and effectors; adenosine receptors are not the major focus of this review (for reviews on adenosine receptors, see Refs. 100, 101, 108, and 133).
On the basis of the differences in rank-potency order profiles of nucleotides and nucleotide analogs, Burnstock and Kennedy suggested a subclassification of P2 receptors into P2X and P2Y subtypes (25). Later studies revealed that P2X and P2Y receptors also differ in their transduction mechanisms (55). P2X receptors appeared to have an intrinsic ion channel that increased the permeability of the plasma membrane to Na+, K+, and Ca2+ (and, possibly, anions), whereas P2Y receptors were traditional G protein-coupled receptors that coupled to heterotrimeric G proteins, phospholipases, and phosphoinositol signaling pathways. Gordon (50) further delineated the existence of P2T and P2Z subtypes of receptors into platelets and mast cells, respectively. These were originally thought to be distinctly different from the P2X and P2Y receptor subtypes. In his seminal review, Gordon also emphasized the concept of biologically relevant release of nucleotides and nucleosides, discussing the sources, effects, and fates of these purinergic agonists. Subsequently, another subtype of P2 receptor that responds to the pyrimidine nucleotide UTP as well as to ATP was identified and termed "P2U receptor" or the pyrimidine receptor. In addition to the P2 receptor subtypes discussed above, receptors that bind adenosine dinucleotide polyphosphates (Ap4A, Ap5A, and Ap6A) were identified and classified as P2D receptors (27, 59). Although possible functions as neurotransmitters or cotransmitters have been suggested, the exact physiological roles of the diadenosine polyphosphates (ApxA) are not yet well established.
![]() |
PURINERGIC RECEPTORS: "METABOTROPIC" P2Y G PROTEIN-COUPLED RECEPTORS AND "IONOTROPIC" ATP-GATED P2X RECEPTOR CHANNELS |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Figure 1 shows the probable
topologies of P2Y and P2X purinergic receptors that we discuss below in
this section.
|
Metabotropic P2Y G Protein-Coupled Receptors
P2Y receptors are G protein-coupled receptors that bind purine and/or pyrimidine nucleotides and their derivatives. This class of receptors includes the cloned mammalian P2Y1, P2Y2, P2Y4, P2Y6, and P2Y11 receptors. The P2Y1 receptor subtype is the original P2Y receptor of the old nomenclature, whereas the P2Y2 receptor is the P2U receptor of the old classification (46). In addition, the pharmacologically characterized, but not yet cloned, P2YADP (or P2T) receptor from platelets is also included in this class (reviewed in Refs. 1, 2, 8, 13, and 82). The cloned P2Y3 receptor represents a species homolog of the P2Y6 receptor. Moreover, the mammalian equivalent of the Xenopus laevis P2Y8 receptor has not yet been cloned (108). Receptors that were initially numbered P2Y5, P2Y7, P2Y9, and P2Y10 were subsequently found not to be receptors for nucleotides (despite their cloning by homology to other P2Y receptor genes) and, therefore, have been deleted from this class (108). Table 1 shows the properties of different subtypes of cloned mammalian P2Y receptors, their agonist rank-potency order, sources of cDNA, and the cell and tissue distribution. P2Y receptor expression in the kidney has been reviewed recently (4, 26, 63).
|
P2Y receptors vary in length from 308 to 379 amino acids, with a
molecular mass of 41-53 kDa in glycosylated form. These receptors have seven transmembrane-spanning -helical hydrophobic regions typical of G protein-coupled receptors. The NH2 terminus of
the receptor proteins is on the extracellular side of the plasma
membrane, whereas the COOH terminus lies on the cytosolic side of the
membrane. The three-dimensional orientation of the transmembrane
domains creates a pocket with positively charged amino acids that
interact with the phosphate groups of the nucleotide ligands
(ATP/ADP/UTP/diadenosine polyphosphates). These proteins are
usually N-glycosylated on their second extracellular loop. Although the
exact roles of glycosylation are not known, it has been suggested that
the carbohydrate moieties stabilize the protein conformation, protect
the receptors from the action of proteases, and modulate receptor
function (for recent and exhaustive reviews, see Refs. 2
and 108).
The different subtypes of P2Y receptors interact with different types
of G proteins through their intracellular loops. These G proteins
usually activate a membrane-bound phosphatidylinositol-specific phospholipase C (PLC), resulting in enhanced formation of inositol 1,4,5-triphosphate (IP3) and mobilization of cytosolic
Ca2+ (Ca
P2Y receptors as a class do not desensitize as readily as other G
protein-coupled receptor subfamilies. However, when they do
desensitize, the mechanism involves either phosphorylation by protein
kinases or uncoupling from the G protein. Suramin is a generalized
antagonist of P2 receptors, except for P2Y4. Presently, there are no specific antagonists available to distinguish the different subtypes of cloned mammalian P2 receptors, a void in the
field that needs to be filled. Nonselective inhibitors such as suramin,
reactive blue 2, and pyrodoxal phosphate-6-azophenyl 2',4'-disulfonic
acid (PPADS) can be used to antagonize either or both of the P2Y and
P2X receptors (16). PPADS blocks the P2Y1
receptors coupled to PLC but not those coupled to the inhibition of
adenylate cyclase. The platelet P2YADP receptor is blocked by 2-propylthio-d--
-difluoromethylene ATP (FPL 66096) and
2-propylthio-
-
-dichloromethylene-d-ATP (ARL). P2Y receptor
expression and function along the nephron will be addressed below (for
recent and comprehensive reviews, see Refs. 2 and 108).
Ionotropic ATP-Gated P2X Receptor Channels
P2X receptors are Ca2+-permeable, nonselective cation channels (reviewed in Refs. 1, 8, 13, 16, and 90). To date, eight isoforms have been identified and cloned: P2X1 through P2X7, and P2XM, the most recent isoform cloned from skeletal muscle (146). P2X7 is thought to be the cDNA that corresponds to the P2Z receptor (139) described by Gordon (50) in hematopoietic cells. Like the P2Y receptors, it is likely that multiple P2X receptor channel subtypes remain to be identified. Each isoform consists of two membrane-spanningP2X receptors have no sequence homology to any other ion channel. In over 300 sequences, obtained by PCR amplifications on epithelial and endothelial cDNA libraries by using degenerate primers and subjected to the basic alignment research tool (BLAST) algorithm by Schwiebert and colleagues (Schwiebert LM, Rice WC, Kudlow BA, Taylor AL, and Schwiebert EM, unpublished observations; Schwiebert EM, Wallace D, King SR, Braunstein GM, Peti-Peterdi J, Hanaoka K, Guay-Woodford LM, Bell PD, Sullivan L, Grantham JJ, and Taylor AL, unpublished observations; 143), only P2X receptor channel subtypes are identified, with no other ion channel family emerging with even low or partial homology. Functionally, P2X receptor channels resemble glutamate receptors, another class of excitatory, ligand-gated receptor ion channels. Like glutamate receptors, P2X receptor channels are opened by a ligand and are permeable to monovalent and divalent cations. Divalent cations permeate as well as block the channel (16). Topologically, they are most similar to the amiloride-sensitive Na+ channels (ENaCs), inwardly rectifying K+ channels (IRKs, ROMKs), and nematode degenerins (16). Like ENaCs and IRKs, P2X receptors form heteromultimeric complexes. With the exception of P2X6, they also form homomultimeric complexes (145). It is reasonable to assume that P2XRs form tetramers, given that the overall topology of the two transmembrane-spanning P2XR channels is similar to that solved by MacKinnon and colleagues (40) for a bacterial two-transmembrane-spanning K+ channel.
Table 2 summarizes the properties of the
cloned P2X receptor channels. There are no specific agonists for the
individual P2X receptor isoforms, but ,
-methylene ATP,
,
-methylene ATP, and benzoyl-benzoyl ATP will stimulate the P2X
receptors without activating the P2Y receptors. Few specific
antagonists exist that do not also inhibit P2Y receptors (see above),
with the exceptions of trinitrophenyl ATP (TNP-ATP) and oxidized ATP
(16). The seven P2X isoforms vary in their ATP-binding
kinetics, conductances, and desensitization. The EC50
values for ATP range from 0.7 to 15 µM for P2X1 through
P2X6 (20). P2X7 requires 300 µM
of ATP to elicit a current (90). The pharmacology of P2X
receptor channels has been reviewed recently in some detail
(99). Single-channel conductances vary significantly from
one subtype to the next. The permeabilities for sodium and potassium
are similar, whereas Ca2+ permeability is much higher.
Despite the higher permeability for divalent cations, Ca2+
and Mg2+ also block the channel. Of the seven isoforms,
only P2X1 and P2X3 desensitize rapidly on
binding to the agonist (20, 30). Interestingly, the most
abundant isoforms expressed in epithelia (Schwiebert EM, Wallace D,
King SR, Braunstein GM, Peti-Peterdi J, Hanaoka K, Guay-Woodford LM,
Bell PD, Sullivan L, Grantham JJ, and Taylor AL, unpublished
observations; 143) and endothelia (Schwiebert LM, Rice WC, Kudlow BA,
Taylor AL, and Schwiebert EM, unpublished observations) are poorly
desensitizing members of the P2X receptor family, P2X4,
P2X5 and, to a lesser extent, P2X2 and
P2X7. P2X receptor channel expression and function along the nephron will be addressed below.
|
![]() |
ATP RELEASE INTO THE EXTRACELLULAR MILIEU ALONG THE NEPHRON: ASSAYS, SOURCES, MECHANISMS, AND STIMULI |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
Extracellular Nucleotides as Agonists
Cytosolic concentrations of ATP are typically 3-5 mM (possibly as high as 10 mM), whereas its extracellular concentration is very low (50, 108, 124, 142). Despite this huge concentration gradient, ATP and other intracellular nucleotides cannot diffuse out through the lipid bilayer of the cells because of their net negative charge. However, highly controlled and regulated release of intracellular nucleotides occurs in many types of cells under physiological conditions. Pathologically, ATP is released from cells during hypoxia, shear stress, loss of cell viability, or cytolysis. Although metabolism of nucleotides in blood alone is slow and inefficient (in vitro half-time of ATP in whole blood is 10 min and in cell-free plasma is 30 min), nucleotides are rapidly cleared while passing through the vascular bed (half-time of nucleotides in perfused lung is ~0.2 s). This is due to the presence of ecto-ATPase, ecto-apyrase, and 5'-nucleotidase activity on the luminal surface of endothelial cells (50, 108, 124, 142). Thus ATP and its metabolites are traditionally thought to act as autocrine or paracrine factors that act within tissues or tissue microenvironments. The concentration of extracellular nucleotides needed to activate the purinergic receptors is very low (0.1-10 µM) compared with their intracellular concentrations. As such, the cells need to release only 0.1% or less of its intracellular ATP pool to trigger autocrine or paracrine ATP signaling (50, 108, 124, 142). Several additional factors must also be considered that may influence the attainment of the effective concentrations of nucleotides in the extracellular milieu. These factors include, but are not limited to, the amount of the nucleotides released, their volume of distribution in extracellular microenvironment, and the presence and activity of ecto-apyrases. Despite these many factors, ATP has been measured in significant concentrations in plasma, bile, and urine (31, 62, 93). Figure 2 integrates the concept of purinergic receptor expression on apical and basolateral membranes, the ATP release mechanisms that may promote ATP release from epithelia (see below), and some examples of epithelial processes that purinergic signaling regulates.
|
ATP Release Mechanisms
There are at least three possible mechanisms for the regulated release of nucleotides under physiological conditions. Exocytosis of ATP-filled vesicles is a major mechanism, especially in platelets, neurons, and neuroendocrine cells, such as adrenal medullary chromaffin cells and mast cells. The dense granules of platelets, mast cells, and chromaffin cells store many agonists; however, they also contain ATP and/or ADP at very high concentrations (mM or higher). Platelet dense granules contain a combined ATP+ADP concentration of ~1 M. Approximately 15% of the dry weight of adrenal medulla is due to ATP. UTP is also released by platelets during platelet aggregation, although the intracellular concentration of UTP is much lower than ATP. As such, ATP and its metabolites are thought to be neurotransmitters or cotransmitters or coagonists with other classic neurotransmitters or histamine (50, 108).Nonexocytotic release of ATP and other nucleotides likely occurs by passive transport mechanisms, using the large gradient for ATP efflux, secretion, or exit. Nonconductive transport of ATP (passively down its large concentration gradient out of the cell) may also be a route of ATP release. Such adenine nucleotide transporters have been studied intensely in mitochondrial membrane (19, 87, 119) and have been idenfied recently in endoplasmic and sarcoplasmic reticulum (53), chromaffin granule ghosts (7), and in rat brain synaptic vesicles (51). Nonconductive, bidirectional transporters exist in the plasma membrane for nucleosides primarily in brain (35); therefore, it cannot be discounted that ATP-specific transporters may exist. Indeed, the cystic fibrosis transmembrane conductance regulator (CFTR), once thought to be an ATP channel, may transport ATP as well as other large organic anions (gluconate, glutathione) at nonconductive rates. By the macropatch-recording method, Linsdell and Hanrahan (85) showed rates of transport of larger organic anions that were on the border between conductive and nonconductive.
ATP-permeable anion channels have been characterized biophysically by
many laboratories. Their relationship to CFTR anion channels has been
controversial (68, 83, 109, 110, 111, 124, 126); however,
recent studies by Engelhart and colleagues (68), Foskett
and co-workers (105, 136), and Schwiebert and colleagues
(17) suggest that CFTR does not conduct ATP itself but
regulates a closely associated anion channel that does conduct ATP.
Anion channels with less selectivity for Cl vs. other
halides or larger anions are prime candidates for putative ATP
channels. These include, but are not limited to, the outwardly rectifying Cl
channel (or ORCC) as well as plasma
membrane forms of the voltage-dependent anion channel (VDAC; also
called "porin") (10, 11, 112, 144). Large-conductance or "maxi" Cl
channels that
resemble mitochondrial VDAC or porin biophysically have been observed
in most if not all cells. Interestingly, the major role of VDAC or
porin in the mitochondrial membrane is the transport of newly
synthesized ATP from the mitochondrion to the cytoplasm. Schwiebert and
co-workers (142) as well as Fitz and colleagues
(113, 114, 115, 149) have shown that ATP release is
stimulated under conditions of hypotonic stress. This release is
immediate (within seconds), and it precedes the time course for
regulatory volume decrease after cell swelling (17, 142). Another elegant example of a paracrine ATP signal involves propagation of Ca2+ waves along monolayers of cells in a gap
junction-independent manner (49, 120). Thus these are key
examples of how ATP works well as a signaling molecule, because it is
released rapidly down its large concentration gradient, it acts in an
autocrine or paracrine manner, it is rapidly degraded to dissipate the
response, and it mediates fast and slow responses, via P2X and P2Y
receptors, respectively (16).
Assays to Detect ATP Release and Signaling
Several laboratories have explored ATP release in the context of excitable and nonexcitable cells and have developed different assays to aid this study. Schwiebert and co-workers (142) developed an assay using a luminometer and the firefly substrate and enzyme luciferin-luciferase by which monolayers of epithelial and endothelial cells, which were tight with fluid and had a resistance level >200Another elegant assay determines ATP release by using luciferin-luciferase attached to the surface of a cell. Luciferase is fused to protein A, a protein that binds to IgG antibodies. Cells are incubated first with an IgG antibody to an extracellular epitope of a specific cell surface protein, then with the protein A-luciferase fusion product. The cells are placed in a luminometer to collect photons as an indicator of localized extracellular ATP release at the membrane surface (9). This assay was utilized on hematopoietic cells, where surface antigens are abundant and well characterized. However, it could be adapted to tissue preparation such as isolated and perfused renal tubules, intestinal crypts, or isolated glands or ducts.
The most recent assay involves atomic force microscopy (AFM) using commercially available AFM tips coated with the myosin subfragment S1, which has a high affinity for ATP and changes shape on ATP hydrolysis. Rather than luciferase-luciferin reagent as the readout, the myosin tips are placed next to a cell, bind the ATP as it is released, and the myosin in the tips responds by changing shape (121). Using this assay, Schneider and co-workers (121) showed that cystic fibrosis airway epithelial cells released little ATP (e.g., the vibration of the AFM probe was minimal); however, when the probe was placed in close proximity to the membranes of cystic fibrosis (CF) cells stably complemented with wild-type CFTR, significant vibration of the probe was observed. Adaptation to isolated tubule preparations has not been performed; however, it is a definite possibility.
Creative indirect assays have also been developed, using the P2X
receptor channels as a functional readout. Hazama et al. (57) developed an assay that determines ATP release from a
single cell. A single PC12 cell, which expresses P2X2, is
lifted while in whole-cell patch configuration and placed near a
pancreatic -cell growing on a coverslip. The pancreatic
-cell is
then stimulated with glucose, while the PC12 cell is studied for
P2X2 current. Current is indicative of the
-cell's
release of ATP (57). This assay is applicable to any cell
or dissected tissue preparation of interest (see below), when
maintained in culture. Moreover, instead of P2X receptor current as an
endpoint, one could load the PC-12 cell with fura 2-acetoxymethyl ester
(AM) and measure Ca2+ influx through P2X2 as a
different fluorescent endpoint (see below). Hollins and Ikeda
(60) used a similar strategy. They transfected rat adrenal
chromaffin cells with P2X1 and then stimulated the cells
with agents that increase exocytosis. Their patch-clamp data showed
that stimulating exocytosis led to P2X1 current, implying that ATP is released in a vesicular manner. Indeed, other laboratories are using these groundbreaking assays or modifying them further with
great success. Moreover, their possible application to the kidney and
to cell or tissue preparations of the nephron was reviewed recently in
more detail (125) and will be expanded on below (see HYPOTHESES AND FUTURE DIRECTIONS).
Putative ATP Release and Release Mechanisms Along the Nephron
Schwiebert and co-workers (Schwiebert EM, Wallace D, King SR, Braunstein GM, Peti-Peterdi J, Hanaoka K, Guay-Woodford LM, Bell PD, Sullivan L, Grantham JJ, and Taylor AL, unpublished observations) have used the bioluminescence detection assay to detect ATP released from primary cultures and cell lines derived from known nephron segments and grown in vitro to determine the major sources of extracellular ATP, the regulation of ATP release, and, ultimately, the mechanisms of ATP release. This assay has also been applied to human vascular endothelial monolayers in primary cultures derived from different blood vessels throughout the vasculature with promising results (Schwiebert LM, Rice WC, Kudlow BA, Taylor AL, and Schwiebert EM, unpublished observations). Pseudopolarized epithelial cell cultures or polarized epithelial cell monolayers were studied in real time for basal and stimulated ATP release and for the sidedness of ATP release, respectively.Figure 3 shows estimated concentrations
of ATP based on parallel standard curves using the same
luciferase-luciferin-containing reagent but with known quantities of
ATP. The proximal tubule is the richest source of ATP under basal
conditions, at concentrations that reach ~1 µM. On stimulation,
this concentration increases; however, maximal ATP release never
exceeds 5-10 µM. A study using human renal primary cultures to
document ATP concentrations measured under basal and hypotonic
conditions with this bioluminescence assay have been published
(151). As other cultures are examined along the nephron,
the amount of ATP release under basal conditions decreases through the
nanomolar range. Collecting duct epithelia release low-nanomolar to
picomolar concentrations of ATP under basal conditions, but collecting
duct cell lines can be induced to release ATP with hypotonicity (a
condition that would be present in diuresis luminally) and with the
Ca2+ agonists ionomycin and thapsigargin. (Schwiebert EM,
Wallace D, King SR, Braunstein GM, Peti-Peterdi J, Hanaoka K,
Guay-Woodford LM, Bell PD, Sullivan L, Grantham JJ, and Taylor AL,
unpublished observations). As a general principle, ATP release into the
apical medium is more robust than basolateral-directed ATP release
(151). Hypotonic challenge stimulates ATP release
immediately and with transient and sustained components of a time
course across the apical and basolateral membranes. Ca2+
agonists stimulate a slow, monophasic rise in ATP release that plateaus
after several minutes. Ca2+ agonists stimulate
apical-directed ATP release, whereas they are without effect on
basolateral release. Present work is focused on studying the regulation
of ATP release as well as ATP release mechanisms in renal epithelial
cell models. Similar work is ongoing in human vascular endothelial
monolayers (Schwiebert LM, Rice WC, Kudlow BA, Taylor AL, and
Schwiebert EM, unpublished observations). The ultimate goal is to adapt
this assay to the study of whole kidney or isolated nephron segments
(see below).
|
Bell and colleagues have used the PC-12 cell biosensor method developed
by Okada and co-workers to detect ATP release across the basolateral
membrane of the macula densa (MD) plaque that lies within the cortical
thick ascending limb (cTAL) juxtaposed to the afferent arterioles and
the glomerulus (Bell PD, personal communication). It is their
hypothesis that ATP is released from macula densa (MD) in response to
cTAL transport of Cl across the MD as an
autocrine/paracrine signal for tubuloglomerular (TG) feedback. With the
use of whole cell patch-clamp recording of PC-12 cells placed near the
basolateral surface of the MD plaque, ATP-gated P2XR current was
detected, indicative of ATP release from the plaque. Similarly, PC-12
cells loaded with the Ca2+-sensitive dye fura 2-AM were
used as a fluorescence biosensor, and an increase in
Ca
transport
inhibitors. To address the mechanisms of ATP release, cells of the MD
plaque were patch-clamped and the ATP permeability of anion channels
recorded in this membrane was examined. The most promising candidate
for an ATP-permeable anion channel was a maxianion channel of
~300 pS that had an ATP conductance of ~100 pS and also had
permeability to other anions such as Cl
and gluconate
(Bell PD, personal communication). Intriguingly, the voltage dependence
of this ATP-permeable channel is similar to the VDAC. As described
above, the role of VDAC or porin in mitochondrial membrane is to
transport newly synthesized ATP out of the mitochondrion into the
cytoplasm (10, 11). If plasma membrane-expressed VDAC
channels exist (112, 144), this may be a principle ATP
release channel. Nevertheless, this is intriguing preliminary work and
is an example of a physiological role for extracellular ATP release and signaling.
Kishore and Knepper incubated freshly microdissected rat inner
medullary collecting duct (IMCD) segments in vitro under physiological conditions at 37°C in Terasaki plates under mineral oil and with oxygenation. They measured the release of ATP by the tubules into the
medium at different time points using the luciferin-luciferase assay
system in a luminometer. Their data showed that that IMCD segments
release ATP at a rate of 2.11 ± 0.29 fmol · mm1 · 30 min
1, with
a range of 1.43-3.3 fmol. This release was confirmed further in
microperfused IMCD, which released ATP into the lumen at a rate of 1.52 fmol · mm
1 · 10 min
1. They
also correlated this result with the total ATP content of the IMCD
segments. The ATP content was 1 log order of magnitude greater than the
rate at which the IMCD released ATP. They concluded that the release
was physiological and did not represent a nonspecific leakage due to
cell death. These preliminary data indicate that ATP can be released by
the collecting duct cells as an autocrine, paracrine, or autoregulatory
mediator (Kishore BK and Knepper MA, unpublished observations).
More indirectly, two laboratories have documented a role for endogenous
ATP release and signaling as a modulator of intracellular signaling set
points. Insel and co-workers (63) have recently published
the conclusion that ATP, released endogenously and continuously by
Madin-Darby canine kidney (MDCK) cells, modulates phosphatidylinositol signaling and turnover as well as cAMP production in MDCK cells. They
showed that removal of the endogenously released ATP with the
ATPase/ADPase apyrase or antagonism of P2 receptors significantly decreased arachidonic acid release as well as cAMP production. They
also observed that vigorous medium changes on MDCK and other cultures
stimulated ATP release mechanically. Similar results have been observed
in fura 2-AM Ca channel activity by endogenous ATP signaling was also
shown in hepatocytes and cholangiocytes by Fitz and co-workers
(114).
Taken together, for purinergic receptor expression to be relevant physiologically (see next section) there must be biologically relevant ATP release occurring in the same tissue, culture, or microenvironment. Not as much attention has been paid to ATP release in the past; however, interest in this aspect is gaining momentum in the purinergic receptor field. It is likely that these fields will merge and synergize to bring the larger picture of autocrine/paracrine extracellular purinergic signaling into focus.
![]() |
SEGMENTAL DISTRIBUTION OF PURINERGIC RECEPTORS ALONG THE NEPHRON EPITHELIUM: PURINERGIC SIGNALING AND REGULATION OF RENAL EPITHELIAL FUNCTION |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
The physiological significance of purinergic signaling in different cell types along the nephron and in specific nephron segments is an emerging and complex area of research. The principles of ATP release, ATP degradation, multiple subtypes and subfamilies of ATP receptors, and their coupling to multiple signal transduction cascades (discussed above) are complex in itself. This complexity is magnified by the 20 or more different cell types along the nephron and the 10 or more discrete nephron segments that are in contact with the tubular fluid or interstitium containing nucleotide or nucleoside agonists. In this section, themes will emerge that presently guide this field and will continue to guide it for some time to come. First, purinergic receptors are expressed abundantly in the kidney on all cell types of the glomerulus, on vascular smooth muscle and endothelial cells of the renal vasculature, and on the renal epithelium in all nephron segments where it has been examined. Second, in a given renal epithelial cell model (MDCK, mIMCD-K2, etc.), multiple P2Y receptors and multiple P2X receptor channels are expressed in the same epithelial cell and, often, in the same membrane domain of the epithelial cell. The reason this redundancy in ATP receptors is needed is an important fundamental question. Redundancy may be needed for an essential physiological process in all cells such as cell volume regulation, or the P2Y receptors and P2X receptor channels may fulfill completely different roles in the same epithelial cell. Third, and most important, how does purinergic signaling as an autocrine and paracrine agonist cascade regulate renal function independently of other agonists and how does it modulate the effects of important endocrine hormones such as vasopressin or angiotensin II or aldosterone? Each of these fundamental questions is just beginning to be addressed.
P2Y Receptors Along the Renal Epithelium
The P2Y2 receptor has been the most often studied P2Y subtype along the nephron. This fact is mainly due to its ubiquitous expression in most, if not, all cells and tissues. Moreover, UTP is the highest affinity agonist for the P2Y2 receptor, formally defined as the P2U receptor, for this reason. Therefore, laboratories could merely test for the expression of the P2Y2 receptor with UTP. This has been complicated by the fact that newly cloned subtypes P2Y4 and, to a lesser extent, P2Y6, are also stimulated by UTP or UDP. As such, all previous studies with UTP cannot be solely attributable to the P2Y2 receptor. P2Y4 involvement cannot be discounted in renal epithelial cell models until its expression in the kidney is ruled out. To our knowledge, expression of P2Y4 has not been investigated in the kidney. In the context of this section, hints in past literature as well as emerging studies will touch on the expression of P2Y1, P2Y6, and P2Y11 in renal epithelial cell models in addition to the extensive work on P2Y2 (Table 3).
|
Kishore et al. (73) mapped P2Y2 receptor
expression along the mammalian nephron. Specific antibodies to the
P2Y2 receptor as well as a gene-specific cDNA probe
confirmed its expression in terminal IMCD as well as its ubiquitous
expression in all regions and nephron segments of the kidney. Apart
from the medullary regions and segments, cortical regions and segments
also showed strong immunohistochemical signal (Kishore BK, unpublished
observations). Of interest, expression of the P2Y2 receptor
on both membrane domains of the IMCD was also found (see Fig.
4), which is similar to the expression of
vasopressin V2 receptor in IMCD (98). This is
both interesting and perplexing, and this issue requires further investigation. The present working model for the regulation of collecting duct water permeability that grew out of the work of Kishore
and co-workers as well as other investigators is shown in Fig.
4. AVP, acting through its V2 receptor and the cAMP
second messenger system, increases osmotic water permeability of the collecting duct apical membrane by translocating the aquaporin-2 (AQP2)
water channel-containing vesicles from a subapical pool to the apical
plasma membrane. The apical membrane is the rate-limiting barrier for
transepithelial water transport, as AQP3 and AQP4 are expressed
constitutively in the basolateral plasma membrane under normal
conditions (for reviews, see Refs. 96 and 97). On the
other hand, the agonist activation of the P2Y2 receptor for
extracellular nucleotides, as well as the endothelin receptor (ETBR) and the EP3 subtype receptor of
prostaglandin E2 (EP3R), antagonize the
AVP-stimulated water transport in collecting duct (18, 43, 58,
71, 79, 116). This antagonism is achieved by virtue of the
cross-talk mechanisms that exist between the two mutually opposing
intracellular signaling pathways in the collecting ducts, as depicted
in Fig. 4 (18, 58, 71, 78, 116).
|
By measuring Ca secretion across monolayers of
mIMCD-K2 cells. Moreover, in the same monolayers, these nucleotide
agonists inhibited Na+ absorption. Taken together, these
studies show that P2Y2 receptor localization and activation
in medullary collecting duct segments have a profound effect on cell
signaling and modulation of salt and water transport across these segments.
Recently, Deetjen and colleagues (36) have performed
elegant fluorescence imaging studies in isolated perfused mouse
cortical collecting duct (CCD) showing that luminal UTP triggers an
increase in Ca
Utilization of heterologous renal epithelial models derived from
kidney, like the distal nephron models in MDCK cells, A6 Xenopus
laevis kidney cells, and LLC-PK1 porcine kidney cells, has provided ideal models in which to study purinergic signaling. As
early as 1979, Simmons (129, 130) showed that exogenous
ATP agonists affect ion transport across MDCK monolayers. Soon after this finding, Simmons and co-workers (131) showed that
extracellular ATP stimulated Cl secretion across MDCK
monolayers. Very recently, they revisited this work in mIMCD-K2 cells
and found that external ATP stimulated Cl
secretion in a
similar manner (14). However, it is Insel and co-workers
(48, 64, 108) who have led the way in this regard with
multiple publications regarding the MDCK model. They have examined
purinergic signaling via phosphoinositide signaling cascades. Arguably,
their work has triggered present work in isolated nephron segments or
in epithelial cell models derived from defined nephron segments. Rather
than discuss their work in detail, Insel (63) has recently
published a review on his work that does better justice to this body of
literature. Zambon and co-workers (155) have shown recently that P2Y1, P2Y2, and
P2Y11 are expressed in MDCK cells, and this laboratory has
cloned the canine forms of these P2Y receptor genes (P. A. Insel,
personal communication). In addition to work by Insel and co-workers,
Kishore, Ecelbarger, and co-workers have done similar work on
nucleotide agonist-dependent signaling in IMCD. Ecelbarger et al.
(42) demonstrated that prior exposure of rat IMCD to
indomethacin, an inhibitor of cyclooxygenase (COX), attenuates
Ca
Indeed, evidence for expression of additional P2Y receptor subtypes is
accumulating. In a few studies highlighted above, P2Y1 expression along with P2Y2 was documented. In the S1
segment of PCT as well as in OMCD, stimulation by a panel of nucleotide
agonists suggested the concomitant expression of P2Y1 and
P2Y2 (29). In a parallel study by the same
laboratory, RT-PCR in cell lines derived from S1 PCT and OMCD showed
expression of P2Y1, P2Y2, and the P2X receptor
channel subtype P2X4 (140). Rank order potency studies of nucleotide agonist effects on Ca channel. Finally, Ishikawa et
al. (65) found that both P2Y1 and
P2Y2 receptors may mediate increased growth rates of rat
renal IMCD cells in vitro. Taken together, P2Y1 alone or
together with P2Y2 may mediate effects on ATP along the
nephron in multiple nephron segments, specifically proximal tubule and
collecting duct.
Very recently, Bailey and co-workers (4) have added a
third P2Y receptor to the list of those subtypes expressed along the
rat renal epithelium. Single-nephron-segment RT-PCR revealed that
expression of P2Y6 is high in PCT and in the thick
ascending limb and thin descending limb of Henle's loop. Expression
was qualitatively lower in OMCD, although RT-PCR is not precisely quantitative. Expression was absent in the thin ascending limb of Henle
as well as in IMCD. Functional expression was assessed by using
basolateral perfusion of the selective nucleotide UDP, which stimulates
P2Y6 receptors more readily than it does other subtypes.
UDP-triggered increases in Ca
P2X Receptor Channels Along the Renal Epithelium
Only as recently as 1998 has the expression of P2X receptor channels been assessed in renal epithelial cell models. Filipovic et al. (47) found functional and molecular evidence for a P2X1-like receptor channel in the heterologous renal epithelial cell model LLC-PK1 (47). As mentioned above, RT-PCR in cell lines derived from S1 PCT and OMCD showed expression of the P2X receptor channel subtype P2X4 (140). In 1999, McCoy et al. (92) showed that P2X3 and P2X4 subtypes are expressed specifically in mIMCD-K2 cells and that P2X receptor channels stimulate ClRecently, Schwiebert and co-workers (Schwiebert EM, Wallace D,
King SR, Braunstein GM, Peti-Peterdi J, Hanaoka K, Guay-Woodford LM,
Bell PD, Sullivan L, Grantham JJ, and Taylor AL, unpublished observations) have attempted to map P2X receptor channel expression in
cell line and primary cultures derived from defined nephron segments.
Degenerate RT-PCR with primers that amplify all seven isoforms of the
P2X receptor gene family was performed. Verification of this work is
ongoing with subtype-specific antibodies. The amplified product was
subcloned for sequencing, and 12-32 white bacterial colonies
bearing a PCR product insert were sequenced for each epithelial cell
mRNA sample. Each sequence derived from each white bacterial colony was
subjected to the BLAST algorithm to define and confirm the identify of
the P2X receptor subtype found (3). In human mixed renal
epithelial primary cultures, two different proximal tubule primary
cultures, in human mesangial cell primary cultures, and in human
autosomal dominant polycystic kidney disease (ADPKD) primary cultures,
P2X4 and P2X5 are expressed abundantly to the
exclusion of all other isoforms. Only human renal mixed epithelial and
human proximal tubule primary cultures expressed a third isoform,
P2X7, and only with low incidence. In the RCCT-28A cell
line, a model of A-type intercalated cells, P2X5 was the
predominant subtype expressed, whereas, in mIMCD-K2 cells,
P2X3 and P2X4 were expressed at equal incidence
to the exclusion of other isoforms. In collecting duct primary cultures from wild-type mice and the cpk autosomal recessive PKD
mouse, P2X3 was the most abundant sequence found, whereas a
smattering of P2X1, P2X2, and P2X4
was also found in much lower incidence. These recent and novel findings
are shown in a schematic fashion in Fig.
5. They need to be confirmed by
biochemical methods as well as single-nephron RT-PCR with degenerate
primers and with primers specific to each subtype; however, they agree
with studies in epithelial cell models from other tissues as well as
the early studies described above. Taken together, these results
suggest that multiple P2X receptor channels, in addition to multiple
P2Y receptors, are poised to receive autocrine/paracrine nucleotide agonist signals as they travel along the renal epithelium.
|
Pathophysiological Paradigms in Which Purinergic Signaling May Be Beneficial or Detrimental
The pathophysiological significance of purinergic signaling in different cell types along the nephron and in specific nephron segments is also being appreciated.Hypertension. One major area in which purinergic signaling may have import is that of hypertension. Because extracellular purinergic signaling modulates NaCl and water handling along the nephron, this signaling could affect certain hypertensive states. Moreover, because salt and water reabsorption is inhibited by extracellular ATP in the tubular lumen or apical environment, delivery of ATP as a therapeutic agent could lower salt and water reabsorption, blood volume, and blood pressure. Certainly, studies by Kishore et al. (71) as well as McCoy et al. (92) agree and underscore this hypothesis. On the other hand, Churchill and Ellis (32) demonstrated that, in rat renal cortical slice preparation, ATP (100-500 µM) and its analogs stimulate renin secretion in a concentration-dependent manner. They also demonstrated that this effect is mediated by a subtype of P2Y receptor via nitric acid (32). Thus it is likely that the multiple purinergic mechanisms may play a role in blood pressure regulation through kidney.
Water balance. Kishore and co-workers (72) have extended their work on the involvement of purinergic signaling in rat models of hypo- and hypervolemia and polyuria of ischemic-reperfusion injury. Preliminary data indicate that the abundance of the P2Y2 purinoceptor protein in the renal medulla is increased significantly in the hydrated state (hypervolemia) compared with dehydrated state (hypovolemia). Furthermore, there is an apparent shift in the subcellular localization of this protein in the medullary collecting duct cells, as revealed by immunoperoxidase labeling with P2Y2 receptor-specific antibody. In the hypovolemic condition, a predominantly basal labeling is seen as opposed to a predominantly apical labeling in the hypervolemic state (72). Interestingly, Kishore and associates (74) also observed that the distribution of immunoreactive cytosolic phospholipase A2 (cPLA2) in the medullary collecting duct cells is also altered under these conditions, with an intense labeling for cPLA2 on the basal aspect of the cells. These preliminary observations, when established further by more extensive studies, will underscore the importance of the association between the purinergic signaling and arachidonic acid metabolism in the long-term conditioning of collecting duct water permeability and thus offer a basis for the vasopressin-independent regulatory mechanisms. Furthermore, in a preliminary study Kishore and colleagues (76) observed that, in the renal medulla of rats during the reperfusion phase after bilateral renal pedicle clamping, P2Y2 purinoceptor mRNA was markedly increased with a concomitant decrease in AQP2 water channel mRNA. Because it has been well documented by Fernandez-Llama and co-workers (45) that decreased aquaporin protein abundance in collecting duct cells is a contributing factor in the increased urine flow seen in moderate postischemic acute renal failure, the preliminary observations of Kishore et al. (76) on P2Y2 expression suggest the probable involvement of this receptor in the diuretic condition of ischemic reperfusion injury (IRI). Moreover, release of ATP and adenosine in other tissues during IRI has been shown and could play additional modulatory roles beyond regulation of water balance by the medulla.
Tubuloglomerular feedback. Studies by Bell and colleagues have shown that ATP signaling in the interstitium between the macula densa plaque in the cTAL and the afferent arteriole-glomerulus complex may aid in tubuloglomerular feedback mechanisms that regulate renal blood flow and glomerular filtration rate (Bell PD, personal communication). In pathophysiological states where these parameters change dramatically, autocrine/paracrine ATP signaling may be affected.
PKD.
Alternatively, in recessive and dominant forms of PKD, extracellular
ATP signaling could prove detrimental. Wilson et al. (151)
showed that human PKD epithelial monolayers release as much or more ATP
than do normal epithelial controls. They also showed that a subset of
human PKD cyst fluid samples had nanomolar to micromolar ATP
concentrations (some as high as 10 µM ATP). All samples had
measurable ATP levels above background, suggesting that ATP release
into the PKD cyst lumen and extracellular ATP signaling within the cyst
occur in vivo. They have expanded this data set into other
PKD models more recently (Schwiebert EM, Wallace D, King SR, Braunstein
GM, Peti-Peterdi J, Hanaoka K, Guay-Woodford LM, Bell PD, Sullivan L,
Grantham JJ, and Taylor AL, unpublished observations). Figure
6 shows our general working hypothesis
concerning the possible detrimental impact of extracellular purinergic
signaling. ADPKD cysts, as fluid-filled structures or spheres
lined by a single layer of epithelial cells, present a physiological
enigma. Ion and fluid transport become encapsulated, especially
secretory transport into the cyst. As such, any secretagogue may be
detrimental. Indeed, growth factors are released and trapped with the
cyst lumen where they interact with growth factor receptors,
creating a vicious autocrine/paracrine growth loop (95,
150). Purinergic agonists are mitogens or comitogens with growth
factors in some cell models (44, 49, 103, 106, 120, 123,
148).
|
![]() |
LESSONS FROM THE STUDY OF EPITHELIA DERIVED FROM OTHER TISSUES |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
A multitude of laboratories have studied purinergic agonists and
their effects of epithelial cell models. In particular, purinergic agonists were among a large panel of agonists tested for their ability
to stimulate Cl and fluid secretion from cystic fibrosis
(CF) tissues and epithelial cell models from the lung and airways and
from the gastrointestinal (GI) system. In CF, Cl
and
fluid secretion are lacking, whereas sodium absorption is augmented.
Purinergic agonists were among the few agonists tested that were
successful in correcting this defective NaCl transport in CF cells
(77, 80, 134, 135). As such, UTP and UTP analogs are being
developed to target the P2Y2 receptor as well as the P2Y4 and P2Y6 receptors to stimulate
Cl
and fluid secretion in CF airways (77, 80, 126,
134, 135). Hwang and colleagues (61) demonstrated
that the addition of purinergic agonists to the apical or basolateral
membrane of rat primary tracheal epithelial cell cultures stimulates
Cl
secretion as a downstream result of activating
purinergic receptors. This study also concluded that the P2Y receptors
stimulated were different on the apical side vs. the basolateral side.
Paradiso and colleagues (104) showed that purinergic
agonists increase Ca
secretion in rabbit tracheal
epithelium, whereas they also observed that ATP suppressed sodium
absorption. This result was confirmed by Devor and Pilewski
(38), who showed that UTP and ATP inhibited sodium
absorption in human non-CF and CF airway epithelium. Taylor et al.
(143) described how P2X receptor activation stimulated Cl
secretion across the nasal epithelium of anesthetized
mice, as well as across mouse and human primary epithelial cell
monolayers. In IMCD monolayers, multiple P2Y and P2X receptors have
been found (92). In the MDCK kidney epithelial model,
Insel (63) found multiple P2Y receptors (P2X receptor
expression in MDCK models has not been examined). Luo and colleagues
(89) also found that multiple P2Y and P2X receptors were
expressed on the luminal and serosal membranes of pancreatic duct,
where they stimulated Cl
secretion and increased
Ca
Collectively, these studies also describe ATP stimulation of
Cl secretion and inhibition of sodium absorption across
airway and kidney epithelia. Because Cl
secretion is
lacking and sodium absorption is elevated in CF, and depending whether
P2Y and/or P2X receptors were targeted with agonists, purinergic
agonist therapy may serve to correct abnormal handling of salt and
water by the respiratory epithelium. Figure 7 shows a working hypothesis concerning
what may happen as a consequence of a lack of extracellular ATP
signaling. Schwiebert and co-workers as well as other laboratories
(68, 121, 136, 143) have shown that ATP release and
signaling are lost in the apical medium bathing CF airway epithelium.
CFTR regulates this process via multiple mechanisms (68, 105,
121, 126, 136, 143), by virtue of its ability to be a
conductance regulator and a regulator of other cellular processes (see
Fig. 7). Extracellular ATP release and signaling also control cell
volume regulation (17, 56, 149). CFTR potentiates this
process as well. In either case, however, CFTR does not conduct ATP as
an anion itself (17, 56, 83, 109, 110, 136). That does not
mean that it may transport ATP at nonconductive rates or regulate
channel-mediated, transporter-mediated, or exocytic mechanisms of ATP
release (84, 111). Under hypotonic stress, cells swell,
release more ATP, and return to their normal cell volume. In the
presence of ATP scavengers (hexokinase, apyrase) or purinergic receptor
antagonists (suramin, reactive blue 2), cells are not capable of
regulatory volume decrease. Additionally, under isotonic conditions,
ATP scavengers cause cells to swell spontaneously, whereas ATP agonists
like ATP
S cause cells to shrink. These data demonstrate the
important role ATP and purinergic receptors play in maintaining cell
volume. Indeed, given the altered ion transport in CFTR mutant cells,
it has been shown that the airway surface liquid (ASL) salt
content and/or volume bathing CF airway epithelia is abnormal compared
with non-CF epithelia. With the use of two different techniques, the
two most recently published articles on ASL have demonstrated that
there is a decreased ASL volume in CF airways that leads to the
inability to clear mucous. The two groups disagree on the salt
composition of the ASL (52, 91, 132, 156). This phenotype
may reflect a more fundamental abnormality in cell volume regulation of
the CF airway epithelium (Fig. 7).
|
To further enhance the important biological role of ATP in CF, Korngreen, Priel, and colleagues (80, 141) demonstrated that ATP increases ciliary beat frequency in freshly isolated rabbit ciliated airway epithelia, whereas Wong and Yeates (153) described the same in vivo on the tracheal lumen of anesthetized dogs. Together, these studies illustrate the importance of extracellular ATP and luminal purinergic receptors in mucus clearance along the airway. An impaired process in CF, enhancement of mucociliary clearance by purinergic agonist therapy would only benefit this disease phenotype as well.
![]() |
HYPOTHESES AND FUTURE DIRECTIONS |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
As this review underscores, the elements of ATP release, ATP receptors, and nucleotide-regulated functions along the nephron are in place for autocrine/paracrine signaling along the renal epithelium. Because the tubular lumen of the nephron is a closed system with regard to ATP filtered at the glomerulus or released into the lumen, this environment is ideal for extracellular purinergic signaling. The complexities of multiple ATP release mechanisms and multiple types of P2Y and P2X receptors expressed by a given renal epithelial cell along the nephron have been delineated by this review. Moreover, ATP release mechanisms and ATP receptors are also present in glomerular mesangium, vascular endothelium, and vascular smooth muscle surrounding renal vessels. This arrangement could allow elaborate purinergic cross-talk among renal nephron, glomerulus, and vasculature that tightly controls kidney function.
Challenges to renal investigators involve simplifying the system to an isolated nephron segment or epithelial cell derived from that nephron segment to dissect out ATP release mechanisms that are expressed. For ATP release to be relevant biologically, ATP receptors must be expressed. They also need to be defined in a given renal epithelial model or in a tubule preparation. Finally, the precise roles of P2Y receptors and P2X receptor channels need definition. It is not as simple as the notion that P2Y receptors expressed luminally and P2X receptors expressed basolaterally. There are mixtures in each membrane of a renal epithelial cell model. Why does the nephron need so many purinergic receptors? This is an open question that requires an answer.
Finally, emerging ATP agonist detection technology needs to be applied
to the kidney. Figure 8 shows some
hypothetical applications of a luciferase-luciferin reagent in
solution, luciferase conjugated to protein A, and an atomic force
microscopy (AFM) probe coated with myosin fragments. Each can be
applied to the kidney, in the right hands and with some ingenuity. The
PC-12 biosensor method has already been applied to the macula densa by
Bell and co-workers, as described above. It is the hope of the authors
of this review that more technology, hard work, and ingenuity be
applied to the concept of purinergic signaling along the nephron.
|
![]() |
ACKNOWLEDGEMENTS |
---|
The authors thank Drs. Mark Knepper, Jim Schafer, and Darwin Bell for critical reading of the manuscript and for helpful suggestions.
![]() |
FOOTNOTES |
---|
E. M. Schwiebert has been supported by New Investigator grants from the American Heart Association [AHA; Southern Research Consortium], the Polycystic Kidney Research Foundation, and by the Cystic Fibrosis Foundation (CFF). E. M. Schwiebert is presently supported by Research Grants from the AHA (SRC) and from the CFF as well as National Institutes of Health Grants R01-DK-54367-01A2 and R01-HL-63934-01. Earlier work of B. K. Kishore was performed in collaboration with Chung-Lin Chou, Mark A. Knepper, and Søren Nielsen at the National Heart, Lung, and Blood Institute (NHLBI; Bethesda, MD) and was supported by the intramural budget of the NHLBI, where he was a Visiting Fellow. Recent studies of B. K. Kishore are being performed in collaboration with Anil G. Menon, Carissa M. Krane, Hamid Rabb, Manoocher Soleimani, Leslie Myatt, Diane Brockman, and Rashmi Sahay at the University of Cincinnati (UC) and are supported by academic development funds from the UC and Dialysis Clinic, Inc. (Cincinnati, OH).
Address for reprint requests and other correspondence: E. M. Schwiebert, Univ. of Alabama at Birmingham, MCLM 740, 1918 Univ. Blvd., Birmingham, AL 35294-0005 (E-mail: eschwiebert{at}physiology.uab.edu).
![]() |
REFERENCES |
---|
![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
---|
1.
Abbracchio, MP,
and
Burnstock G.
Purinoceptors: are there families of P2X and P2Y purinoceptors.
Pharmacol Ther
64:
445-475,
1994[ISI][Medline].
2.
Abbracchio, MP,
and
Burnstock G.
Purinergic signaling: pathophysioligcal roles.
Jpn J Pharmacol
78:
113-145,
1998[ISI][Medline].
3.
Altschul, SF,
Gish W,
Miller W,
Myers EW,
and
Lipman DJ.
Basic local alignment search tool.
J Mol Biol
5:
403-410,
1990.
4.
Bailey, MA,
Hillman KA,
and
Unwin RJ.
P2 receptors in the kidney.
J Auton Nerv Syst
81:
264-270,
2000[ISI][Medline].
5.
Bailey, MA,
Imbert-Teboul M,
Turner C,
Marsy S,
Srai K,
Burnstock G,
and
Unwin RJ.
Axial distribution and characterization of basolateral P2Y receptors along the rat renal tubule.
Kidney Int
58:
1893-1901,
2000[ISI][Medline].
6.
Banderali, U,
Brochiero E,
Lindenthal S,
Raschi C,
Bogliolo S,
and
Ehrenfeld J.
Control of apical membrane chloride permeability in the renal A6 cell line by nucleotides.
J Physiol (Lond)
519:
737-751,
1999
7.
Bankston, LA,
and
Guidotti G.
Characterization of ATP transport into chromaffin granule ghosts. Synergy of ATP and serotonin accumulation in chromaffin granule ghosts.
J Biol Chem
271:
17132-17138,
1996
8.
Barnard, EA,
Simon J,
and
Webb TE.
Nucleotide receptors in the nervous system. An abundant component using diverse transduction mechanisms.
Mol Neurobiol
15:
103-129,
1997[ISI][Medline].
9.
Beigi, R,
Kobatake E,
Masuo A,
and
Dubyak GR.
Detection of local ATP release from activated platelets using cell surface-attached firefly luciferase.
Am J Physiol Cell Physiol
276:
C267-C278,
1999
10.
Benz, R.
Biophysical properties of porin pores from mitochondrial outer membrane of eukaryotic cells.
Experientia
46:
131-137,
1990[ISI][Medline].
11.
Benz, R,
Wojtczak L,
Bosch W,
and
Brdiczka D.
Inhibition of adenine nucleotide transport through the mitochondrial porin by a synthetic polyanion.
FEBS Lett
231:
75-80,
1988[ISI][Medline].
12.
Bidet, M,
De Renzis G,
Martial S,
Rubera I,
Tauc M,
and
Poujeol P.
Extracellular ATP increases [Ca2+]i in distal tubule cells. Evidence for a P2Y2 purinoceptor.
Am J Physiol Renal Physiol
279:
F92-F101,
2000
13.
Boarder, MR,
Weisman GA,
Turner JT,
and
Wilkinson GF.
G protein-coupled P2 purinoceptors: from molecular biology to functional responses.
Trends Pharmacol Sci
16:
133-139,
1995[ISI][Medline].
14.
Boese, SH,
Glanville M,
Aziz O,
Gray MA,
and
Simmons NL.
Calcium and cAMP-activated chloride conductances mediate chloride secretion in a mouse renal inner medullary collecting duct cell line.
J Physiol (Lond)
523:
325-338,
2000
15.
Bouyer, P,
Palais M,
Cougnon M,
Hulin P,
Anagnostopoulos T,
and
Planelles G.
Extracellular ATP raises cytosolic calcium and activates a basolateral chloride conductance in Necturus proximal tubule.
J Physiol (Lond)
510:
535-548,
1998
16.
Brake, AJ,
and
Julius D.
Signaling by extracellular nucleotides.
Annu Rev Cell Dev Biol
12:
519-541,
1996[ISI][Medline].
17.
Braunstein, GM,
Roman RM,
Clancy JP,
Kudlow BA,
Taylor AL,
Shylonsky GV.,
Jovov B,
Ismailov,
Peter K, II,
Jilling T,
Collawn JF,
Benos DJ,
Schwiebert LM,
Fitz JG,
and
Schwiebert EM.
CFTR facilitates ATP release by stimulating a separate ATP release channel for autocrine regulation of cell volume.
J Biol Chem.
276:
6621-6630,
2001
18.
Breyer, MD,
and
Breyer RM.
Prostaglanding E receptors and the kidney.
Am J Physiol Renal Physiol
279:
F12-F23,
2000
19.
Brustovetsky, N,
Becker A,
Klingenberg M,
and
Bamberg E.
Electrical currents associated with nucleotide transport by the reconstituted mitochondrial ADP/ATP carrier.
Proc Natl Acad Sci USA
93:
664-668,
1996
20.
Buell, G,
Collo G,
and
Rassendren F.
P2X receptors: an emerging channel family.
Eur J Neurosci
8:
2221-2228,
1996[ISI][Medline].
21.
Burnstock, G.
Purinergic nerves.
Pharmacol. Rev
24:
509-581,
1972[ISI][Medline].
22.
Burnstock, G.
Purinergic nerves and receptors.
Prog Biochem Pharmacol
16:
141-154,
1980[Medline].
23.
Burnstock, G.
Purinergic receptors in the heart.
Circ Res
46:
I175-I182,
1980[ISI][Medline].
24.
Burnstock, G,
Campbell G,
Stachell D,
and
Smythe A.
Evidence that adenosine triphosphate or a related nucleotide is the transmitter substance released by non-adrenergic inhibitory nerves in the gut.
Br J Pharmacol
40:
668-688,
1970[ISI][Medline].
25.
Burnstock, G,
and
Kennedy C.
Is there a basis for distinguishing two types of P2-purinoceptors?
Gen Pharmacol
16:
433-440,
1985[Medline].
26.
Burnstock, G,
and
Williams M.
P2 purinergic receptors: modulation of cell function and therapeutic potential.
J Pharmacol Exp Ther
295:
862-869,
2000
27.
Castro, E,
Pinter J,
and
Miras-Portugal MT.
Ca2+-stores mobilization by diadenosine triphosphate Ap4A, through a putative P2Y purinoceptor in adrenal chromaffin cells.
Br J Pharmacol
106:
833-837,
1992[Abstract].
28.
Cha, SH,
Jung KY,
and
Endou H.
Effect of P2Y-purinoceptor stimulation on renal gluconeogenesis in rats.
Biochem Res Commu
211:
454-461,
1995[ISI].
29.
Cha, SH,
Sekine T,
and
Endou H.
P2 purinoceptor localization along rat nephron and evidence suggesting existence of subtypes P2Y1 and P2Y2.
Am J Physiol Renal Physiol
274:
F1006-F1014,
1998
30.
Chan, CM,
Unwin RJ,
Bardini M,
Oglesby IB,
Ford AP,
Townsend-Nicholson A,
and
Burnstock G.
Localization of P2X1 purinoceptors by autoradiography and immunocytochemistry in rat kidneys.
Am J Physiol Renal Physiol
274:
F799-F804,
1998
31.
Chan, CM,
Unwin RJ,
and
Burnstock G.
Potential functional roles for extracellular ATP in kidney and urinary tract.
Exp Nephrol
6:
200-207,
1998[ISI][Medline].
32.
Churchill, PC,
and
Ellis VR.
Purinergic P2Y receptors stimulate renin secretion by rat renal cortical slices.
J Pharmacol Exp Ther
266:
160-163,
1993[Abstract].
33.
Cornfield, LJ,
Hu S,
Hurt SD,
and
Sills MA.
[3H]2-phenylamino-adenosine ([3H]CV 1808) labels a novel adenosine receptor in rat brain.
J Pharmacol Exp Ther
263:
552-561,
1992[Abstract].
34.
Cuffe, JE,
Bielfeld-Ackermann A,
Thomas J,
Leipziger J,
and
Korbmacher C.
ATP stimulates Cl secretion and reduces amiloride-sensitive Na+ absorption in M-1 mouse cortical collecting duct cells.
J Physiol (Lond)
524:
77-90,
2000
35.
Dalziel, HH,
and
Westfall DP.
Receptors for adenine nucleotides and nucleosides: subclassification, distribution, and molecular characterization.
Pharmacol Rev
46:
449-466,
1994[ISI][Medline].
36.
Deetjen, P,
Thomas J,
Lehrmann H,
Kim SJ,
and
Leipziger J.
The luminal P2Y receptor in the isolated perfused mouse cortical collecting duct.
J Am Soc Nephrol
11:
1798-1806,
2000
37.
Degenring, FH,
Rubio R,
and
Berne RM.
Adenine nucleotide metabolism during cardiac hypertrophy and ischemia in rats.
J Mol Cell Cardiol
7:
105-113,
1975[ISI][Medline].
38.
Devor, DC,
and
Pilewski JM.
UTP inhibits sodium absorption in wild-type and delF508 CFTR-expressing human bronchial epithelia.
Am J Physiol Cell Physiol
276:
C827-C837,
1999
39.
Dillingham, MA,
and
Anderson RJ.
Purinergic regulation of basal and arginine vasopressin-stimulated hydraulic conductivity in rabbit cortical collecting tubule.
J Membr Biol
88:
277-281,
1985[ISI][Medline].
40.
Doyle, DA,
Morais Cabral J,
Pfuetzner RA,
Kuo A,
Gulbis JM,
Cohen SL,
Chait BT,
and
MacKinnon R.
The structure of the potassium channel: molecular basis of K+ conduction and selectivity.
Science
280:
69-77,
1998
41.
Drury, AN,
and
Szent-Gyorgyi A.
The physiological activity of adenine compounds with special reference to their action upon the mammalian heart.
J Physiol (Lond)
68:
213-237,
1929.
42.
Ecelbarger, CA,
Maeda CC,
Gibson Y,
and
Knepper MA.
Extracellular ATP increases intracellular calcium in rat collecting duct via a nucleotide receptor.
Am J Physiol Renal Fluid Electrolyte Physiol
267:
F998-F1006,
1994
43.
Edwards, RM,
Stack EJ,
Pullen M,
and
Nambi P.
Endothelin inhibits vasopressin action in rat inner medullary collecting duct via the ETB receptor.
J Pharmacol Exp Ther
267:
1028-1033,
1993[Abstract].
44.
Erlinge, D.
Extracellular ATP: A growth factor for vascular smooth muscle cells.
Gen Pharmacol
31:
1-8,
1998[Medline].
45.
Fernandez-Llama, P,
Andrews P,
Turner R,
Saggi S,
Dimari J,
Kwon TH,
Nielsen S,
Safirstein S,
and
Knepper MA.
Decreased abundance of collecting duct aquaporins in post-ischemic renal failure in rats.
J Am Soc Nephrol
10:
1658-1668,
1999
46.
Fredholm, BB,
Burnstock G,
Harden TK,
and
Spedding M.
Receptor nomenclature.
Drug Dev Res
39:
461-466,
1966.
47.
Filipovic, DM,
Adebanjo OA,
Zaidi M,
and
Reeves WB.
Functional and molecular evidence for P2X receptors in LLC-PK1 cells.
Am J Physiol Renal Physiol
274:
F1070-F1077,
1998
48.
Firestein, BL,
Xing M,
Hughes RJ,
Corvera CU,
and
Insel PA.
Heterogeneity of P2U- and P2Y-purinergic receptor regulation of phospholipases in MDCK cells.
Am J Physiol Renal Fluid Electrolyte Physiol
271:
F610-F618,
1996
49.
Frame, MK,
and
de Feijter AW.
Propagation of mechanically induced intercellular calcium waves via gap junctions and ATP receptors in rat liver epithelial cells.
Exp Cell Res
230:
197-207,
1997[ISI][Medline].
50.
Gordon, JL.
Extracellular ATP: effects, sources and fate.
Biochem J
233:
309-319,
1986[ISI][Medline].
51.
Gualix, J,
Pintor J,
and
Miras-Portugal MT.
Characterization of nucleotide transport into rat brain synaptic vesicles.
J Neurochem
73:
1098-1104,
1999[ISI][Medline].
52.
Guggino, WB.
CFTR and the salt controversy.
Cell
96:
607-610,
1999[ISI][Medline].
53.
Guillen, E,
and
Hirschberg CB.
Transport of adenosine triphosphate into endoplasmic reticulum proteoliposomes.
Biochemistry
34:
5472-5476,
1995[ISI][Medline].
54.
Harada, H,
Chan CM,
Loesch A,
Unwin R,
and
Burnstock G.
Induction of proliferation and apoptotic death via P2Y and P2X receptors, respectively, in glomerular mesangial cells.
Kidney Int
57:
949-958,
2000[ISI][Medline].
55.
Harden, TK,
Boyer JL,
and
Nicholas RA.
P2 purinoceptors: subtype-associated signaling responses and structure.
Annu Rev Pharmacol Toxicol
35:
541-579,
1995[ISI][Medline].
56.
Hazama, A,
Fan HT,
Abdullaev I,
Maeno E,
Tanaka S,
Ando-Akatsuka Y,
and
Okada Y.
Swelling-activated, cystic fibrosis transmembrane conductance regulator-augmented ATP release and chloride conductances in murine C127 cells.
J Physiol (Lond)
523:
1-11,
2000
57.
Hazama, A,
Hayashi S,
and
Okada Y.
Cell surface measurement of ATP release from single pancreatic cells using a novel biosensor technique.
Pflügers Arch
437:
31-35,
1998[ISI][Medline].
58.
Hebert, RL,
Jacobson HR,
and
Breyer MD.
PGE2 inhibits AVP-induced water flow in cortical collecting ducts by protein kinase C activation.
Am J Physiol Renal Fluid Electrolyte Physiol
259:
F318-F325,
1990
59.
Hilderman, RH,
Martin M,
Zimmerman JK,
and
Pivorum EB.
Identification of a unique membrane receptor for adenosine 5', 5"=P1, P4-tetraphosphate.
J Biol Chem
266:
6915-6918,
1991
60.
Hollins, B,
and
Ikeda SR.
Heterologous expression of a P2X-purinoceptor in rat chromaffin cells detects vesicular ATP release.
J Neurophysiol
78:
3069-3076,
1997
61.
Hwang, TH,
Schwiebert EM,
and
Guggino WB.
Apical and basolateral ATP stimulate tracheal epithelial chloride secretion via multiple purinergic receptors and signaling pathways.
Am J Physiol Cell Physiol
270:
C1611-C1623,
1996
62.
Inscho, EW,
Mitchell KD,
and
Navar LG.
Extracellular ATP in the regulation of renal microvascular function.
FASEB J
8:
319-328,
1994
63.
Insel PA. ATP- and nucleotide-mediated signaling via multiple P2Y
receptors in renal epithelial cells. Clin Exp Pharmacol Physiol.
In press.
64.
Insel, PA,
Firestein BL,
Xing M,
Post SR,
Jacobson JP,
Balboa MA,
and
Hughes RJ.
P2 purinoceptors utilize multiple signaling pathways in MDCK-D1 cells.
J Auton Pharmacol
16:
311-313,
1996[ISI][Medline].
65.
Ishikawa, S,
Higashiyama M,
Kusaka I,
Saito T,
Nagasaka S,
and
Fukuda S.
Extracellular ATP promotes cellular growth of renal inner medullary collecting duct cells mediated by P2U receptors.
Nephron
76:
208-214,
1997[ISI][Medline].
66.
Ishikawa, S,
Kawasumi M,
Kusaka I,
Komatsu N,
Iwao N,
and
Saito T.
Extracellular ATP promotes cellular growth of glomerular mesangial cells mediated via phospholipase C.
Biochem Biophys Res Commun
202:
234-240,
1994[ISI][Medline].
67.
Iwase, N,
Sasaki T,
Shimura S,
Yamamoto M,
Suzuki S,
and
Sharito K.
ATP-induced chloride secretion with supressed sodium absorption in rabbit tracheal epithelium.
Respir Physiol
107:
173-180,
1997[ISI][Medline].
68.
Jiang, Q,
Mak D,
Devidas S,
Schwiebert EM,
Bragin A,
Zhang Y,
Skach WR,
Guggino WB,
Foskett JK,
and
Engelhardt JF.
CFTR associated ATP release is controlled by a "chloride sensor" within the channel pore.
J Cell Biol
143:
645-657,
1998
69.
Jin, W,
and
Hopfer U.
Purinergic-mediated inhibition of Na+,K+-ATPase in proximal tubule cells: elevated cytosolic calcium is not required.
Am J Physiol Cell Physiol
272:
C1169-C1177,
1997
70.
Katori, M,
and
Berne RM.
Release of adenosine from anoxic hearts. Relationship to coronary flow.
Circ Res
19:
420-425,
1966[ISI][Medline].
71.
Kishore, BK,
Chou CL,
and
Knepper MA.
Extracellular nucleotide receptor inhibits AVP-stimulated water permeability in inner medullary collecting duct.
Am J Physiol Renal Fluid Electrolyte Physiol
269:
F863-F869,
1995
72.
Kishore, BK,
Di Iulio D,
Menon AG,
and
Krane CM.
Altered abundance and subcellular distribution of P2Y2-purinoceptor protein in inner medulla of thirsted and hydrated rats: possible basis for vasopressin-independent regulation of water transport (Abstract).
J Am Soc Nephrol
11:
17A,
2000.
73.
Kishore, BK,
Ginns SM,
Krane CM,
Nielsen S,
and
Knepper MA.
Cellular localization of P2Y2 purinoceptor in rat renal inner medulla and lung.
Am J Physiol Renal Physiol
278:
F43-F51,
2000
74.
Kishore, BK,
Myatt L,
Brockman D,
and
Sahay R.
Differential subcellular localization of cytosolic phospholipase A2 in inner medulla of dehydrated and hydrated rats (Abstract).
J Am Soc Nephrol
11:
433A,
2000.
75.
Kishore, BK,
Myatt L,
DiIulio D,
and
Brockman D.
MAP-kinase-dependent production of prostaglandin E2 by the agonist stimulation of P2Y2-purinoceptor in rat inner medullary collecting duct (Abstract).
J Am Soc Nephrol
11:
433A,
2000.
76.
Kishore, BK,
Wang Z,
Rabb H,
Haq M,
and
Soleimani M.
Upregulation of P2Y2 purinoceptor during ischemic reperfusion injury (IRI): possible relevance to diuresis of IRI (Abstract).
J Am Soc Nephrol
9:
580A,
1998.
77.
Knowles, MR,
Clarke LL,
and
Boucher RC.
Extracellular ATP and UTP induce chloride secretion in nasal epithelia of cystic fibrosis patients and normal subjects in vivo.
Chest
101:
60-63,
1992.
78.
Kohan, DE.
Endothelin: renal tubule synthesis and actions.
Clin Exp Pharmacol Physiol
23:
337-344,
1996[ISI][Medline].
79.
Kohan, DE,
and
Hughes AK.
Autocrine role of endothelin in rat IMCD: inhibition of AVP-induced cAMP accumulation.
Am J Physiol Renal Fluid Electrolyte Physiol
265:
F126-F129,
1993
80.
Korngreen, A,
Ma W,
Priel Z,
and
Silberberg SD.
Extracellular ATP directly gates a cation-selective channel in rabbit airway ciliated epithelial cells.
J Physiol (Lond)
508:
703-720,
1998
81.
Koster, HP,
Hartog A,
van Os CH,
and
Bindels RJ.
Inhibition of Na+ and Ca2+ reabsorption by P2U purinoceptors requires PKC but not Ca2+ signaling.
Am J Physiol Renal Fluid Electrolyte Physiol
270:
F53-F60,
1996
82.
Kunapuli, SP,
and
Daniel JL.
P2 receptor subtypes in the cardiovascular system.
Biochem J
336:
513-23,
1998[ISI][Medline].
83.
Li, C,
Ramjessingh M,
and
Bear CE.
Purified cystic fibrosis transmembrane conductance regulator (CFTR) does not function as an ATP channel.
J Biol Chem
271:
11623-11626,
1996
84.
Linsdell, P,
and
Hanrahan JW.
Adenosine triphosphate-dependent asymmetry of anion permeation in the cystic fibrosis transmembrane conductance regulator chloride channel.
J Gen Physiol
111:
601-614,
1998
85.
Linsdell, P,
and
Hanrahan JW.
Glutathione permeability of CFTR.
Am J Physiol Cell Physiol
275:
C323-C326,
1998
86.
Lodos, C,
Cooper DMF,
and
Wolff J.
Subclasses of external adenosine receptors.
Proc Natl Acad Sci USA
77:
2551-2554,
1980[Abstract].
87.
Lohret, TA,
Murphy RC,
Drgon T,
and
Kinnally KW.
Activity of the mitochondrial multiple conductance channel is independent of the adenine nucleotide translocator.
J Biol Chem
271:
4846-4849,
1996
88.
Lu, M,
MacGregor GG,
Wang W,
and
Giebisch G.
Extracellular ATP inhibits the small-conductance K+ channel on the apical membrane of the cortical collecting duct from mouse kidney.
J Gen Physiol
116:
299-310,
2000
89.
Luo, X,
Zheng W,
Yan M,
Lee MG,
and
Muallem S.
Multiple functional P2X and P2Y receptors in the luminal and basolateral membranes of pancreatic duct cells.
Am J Physiol Cell Physiol
277:
C205-C215,
1999
90.
Mackenzie, AB,
Surprenant A,
and
North RA.
Functional and molecular diversity of purinergic ion channel receptors.
Ann NY Acad Sci
868:
716-729,
1999
91.
Matsui, H,
Grubb BR,
Tarran R,
Randell SH,
Gatzy JT,
Davis CW,
and
Boucher RC.
Evidence for periciliary liquid layer depletion, not abnormal ion composition, in the pathogenesis of cystic fibrosis airway disease.
Cell
95:
1005-1015,
1999[ISI].
92.
McCoy, DE,
Taylor AL,
Kudlow BA,
Karlson KE,
Slattery MJ,
Schwiebert LM,
Schwiebert EM,
and
Stanton BA.
Nucleotides regulate NaCl transport across mIMCD-K2 cells via P2X and P2Y purinergic receptors.
Am J Physiol Renal Physiol
277:
F552-F559,
1999
93.
McGill, JM,
Basavappa S,
Gettys TW,
and
Fitz JG.
Adenosine triphosphate activates ion permeabilities in biliary epithelial cells.
Gastroenterology
107:
236-243,
1994[ISI][Medline].
94.
Middleton, JP,
Mangel AW,
Basavappa S,
and
Fitz JG.
Nucleotide receptors regulate membrane ion transport in renal epithelial cells.
Am J Physiol Renal Fluid Electrolyte Physiol
264:
F867-F873,
1993
95.
Murcia, NS,
Woychik RP,
and
Avner ED.
The molecular biology of polycystic kidney disease.
Pediatr Nephrol
12:
721-726,
1998[ISI][Medline].
96.
Nielsen, S,
Kwon TH,
Christensen BM,
Promeneur D,
Frokiaer J,
and
Marples D.
Physiology and pathophysiology of renal aquaporins.
J Am Soc Nephrol
10:
647-663,
1999
97.
Nielsen, S,
Kwon TH,
Frokiaer J,
and
Knepper MA.
Key roles of renal aquaporins in water balance and water-balance disorders.
News Physiol Sci
15:
136-143,
2000
98.
Nonoguchi, H,
Owada A,
Kobayashi N,
Takyama M,
Terada Y,
Koike J,
Uije K,
Marumo F,
Sakai T,
and
Tomita K.
Immunohistochemical localization of V2 vasopressin receptor along the nephron and functional role of luminal V2 receptor in terminal inner medullary collecting ducts.
J Clin Invest
96:
1768-1778,
1995[ISI][Medline].
99.
North, RA,
and
Suprenant A.
Pharmacology of cloned P2X receptors.
Annu Rev Pharmacol Toxicol
40:
563-580,
2000[ISI][Medline].
100.
Olah, ME,
and
Stiles GL.
Adenosine receptors.
Annu Rev Physiol
54:
211-225,
1992[ISI][Medline].
101.
Olah, ME,
and
Stiles GL.
Adenosine receptor subtypes: characterization and therapeutic regulation.
Annu Rev Pharmacol Toxicol
35:
581-606,
1995[ISI][Medline].
102.
Ostrom, RS,
Gregorian C,
and
Insel PA.
Cellular release of and response to ATP as key determinants of the set-point of signal transduction pathways.
J Biol Chem
275:
11735-11739,
2000
103.
Paller, MS,
Schnaith EJ,
and
Rosenberg ME.
Purinergic receptors mediate cell proliferation and enhanced recovery from renal ischemia by adenosine triphosphate.
J Lab Clin Med
131:
174-183,
1998[ISI][Medline].
104.
Paradiso, AM,
Mason SJ,
Lazarowski ER,
and
Boucher RC.
Membrane-restricted regulation of calcium release and influx in polarized epithelia.
Nature
377:
643-646,
1995[ISI][Medline].
105.
Pasyk, EA,
and
Foskett JK.
Cystic fibrosis transmembrane conductance regulator (CFTR)-associated ATP and adenosine 3'-phosphate 5'phosphosulfate channels in endoplasmic reticulum and plasma membranes.
J Biol Chem
273:
7746-7751,
1997.
106.
Popper, LD,
and
Batra S.
Calcium mobilization and cell proliferation activated by extracellular ATP in human ovarian tumour cells.
Cell Calcium
14:
209-218,
1993[ISI][Medline].
107.
Post, SR,
Rump LC,
Zambon A,
Hughes RJ,
Buda MD,
Jacobson JP,
Kao CC,
and
Insel PA.
ATP activates cAMP production via multiple purinergic receptors in MDCK-D1 epithelial cells. Blockade of an autocrine/paracrine pathway to define receptor preference of an agonist.
J Biol Chem
273:
23093-23097,
1998
108.
Ralevic, V,
and
Burnstock G.
Receptors for purines and pyrimidines.
Pharmacol Rev
50:
413-492,
1998
109.
Reddy, MM,
Quinton PM,
Haws C,
Wine JJ,
Grygorczyk R,
Tabcharani JA,
Hanrahan JW,
Gunderson KL,
and
Kopito RR.
Failure of the cystic fibrosis transmembrane conductance regulator to conduct ATP.
Science
271:
1876-1879,
1996[Abstract].
110.
Reddy, MM,
Quinton PM,
Haws C,
Wine JJ,
Grygorczyk R,
Tabcharani JA,
Hanrahan JW,
Gunderson KL,
and
Kopito RR.
Cystic fibrosis transmembrane conductance regulator and adenosine triphosphate (Abstract).
Science
275:
1325,
1997[ISI].
111.
Reisin, IL,
Prat AG,
Abraham EH,
Amara JF,
Gregory RJ,
Ausiello DA,
and
Cantiello HF.
The cystic fibrosis transmembrane conductance regulator is a dual ATP and chloride channel.
J Biol Chem
269:
20584-20591,
1994
112.
Reymann, S,
Florke H,
Heiden M,
Jakob C,
Stadtmuller U,
Steinacker P,
Lalk VE,
Pardowitz I,
and
Thinnes FP.
Further evidence for multitopological localization of mammalian porin (VDAC) in the plasmalemma forming part of a chloride channel complex affected in cystic fibrosis and encephalomyopathy.
Biochem Mol Med
54:
75-87,
1995[ISI][Medline].
113.
Roman, RM,
Feranchak AP,
Davison A,
Schwiebert EM,
and
Fitz JG.
Evidence for Gd3+ as an inhibitor of membrane ATP permeability and purinergic signaling.
Am J Physiol Gastrointest Liver Physiol
277:
G1222-G1230,
1999
114.
Roman, RM,
Feranchak AP,
Salter KD,
Wang Y,
and
Fitz JG.
Endogenous ATP regulates Cl secretion in cultured human and rat biliary epithelial cells.
Am J Physiol Gastrointest Liver Physiol
276:
G1391-G1400,
1999
115.
Roman RM, Lomri N, Braunstein GM, Feranchak AP, Simeoni L, L, Davison
AK, Mechetner EB, Schwiebert EM, and Fitz JG. Multidrug resistance
P glycoproteins regulate cellular ATP permeability. J Membr Biol.
In press.
116.
Rouse, D,
Leite M,
and
Suki WN.
ATP inhibits the hydrosmotic effect of AVP in rabbit CCT: evidence for a nucleotide P2U receptor.
Am J Physiol Renal Fluid Electrolyte Physiol
267:
F289-F295,
1994
117.
Rubera, I,
Tauc M,
Bidet M,
Verheecke-Mauze C,
De Renzis G,
Poujeol C,
Cuiller B,
and
Poujeol P.
Extracellular ATP increases [Ca2+]i in distal tubule cells. II. Activation of a Ca2+-dependent Cl conductance.
Am J Physiol Renal Physiol
279:
F102-F111,
2000
118.
Rubio, R,
and
Berne RM.
Release of adenosine by the normal myocardium in dogs and its relationship to the regulation of coronary resistance.
Circ Res
25:
407-415,
1969[ISI][Medline].
119.
Ruck, A,
Dolder M,
Walliman T,
and
Brdiczka D.
Reconstituted adenine nucleotide translocase forms a channel for small molecules comparable to the mitohondrial permeability transition pore.
FEBS Lett
426:
97-101,
1998[ISI][Medline].
120.
Sauer, H,
Hescheler J,
and
Wartenberg M.
Mechanical strain-induced calcium waves in human prostate cancer cells are propagated via ATP release and purinergic receptor activation.
Am J Physiol Cell Physiol
279:
C295-C307,
2000
121.
Schneider, SW,
Egan ME,
Jena BP,
Guggino WB,
Oberleitner H,
and
Geibel JP.
Continuous detection of extracellular ATP on living cells by using atomic force microscopy.
Proc Natl Acad Sci USA
96:
12180-12185,
1999
122.
Schulze-Lohoff, E,
Hugo C,
Rost S,
Arnold S,
Gruber A,
Brune B,
and
Sterzel RB.
Extracellular ATP causes apoptosis and necrosis of cultured mesangial cells via P2Z/P2X7 receptors.
Am J Physiol Renal Physiol
275:
F962-F967,
1998
123.
Schulze-Lohoff, E,
Zanner S,
Ogilvie A,
and
Sterzel RB.
Extracellular ATP stimulates proliferation of cultured mesangial cells via P2-purinergic receptors.
Am J Physiol Renal Fluid Electrolyte Physiol
263:
F374-F383,
1992
124.
Schwiebert, EM.
ABC transporter-faciliated ATP transport.
Am J Physiol Cell Physiol
276:
C1-C8,
1999
125.
Schwiebert EM. ATP release mechanisms, ATP receptors, and
purinergic signaling along the nephron. Clin Exp Pharmacol
Physiol. In press.
126.
Schwiebert, EM,
Egan ME,
Hwang TH,
Fulmer SB,
Allen SS,
Cutting GR,
and
Guggino WB.
CFTR regulates outwardly rectifying chloride channels through an autocrine mechanism involving ATP.
Cell
81:
1063-1073,
1995[ISI][Medline].
129.
Simmons, NL.
Identification of a purine (P2) receptor linked to ion transport in a cultured renal (MDCK) epithelium.
Br J Pharmacol
73:
379-384,
1981[Abstract].
130.
Simmons, NL.
Ion transport in high-resistance dog-kidney cell monolayers: the effect of adenosine 5'-triphosphate.
J Physiol (Lond)
290:
28P-29P,
1979[Medline].
131.
Simmons, NL.
Stimulation of chloride secretion by exogenous ATP in cultured MDCK epithelial monolayers.
Biochim Biophys Acta
646:
231-242,
1981[ISI][Medline].
132.
Smith, JJ,
Travis SM,
Greenberg EP,
and
Welsh MJ.
Cystic fibrosis airway epithelia fail to kill bacteria because of abnormal airway surface fluid.
Cell
85:
229-236,
1996[ISI][Medline].
133.
Spielman, WS,
and
Arend LJ.
Adenosine receptors and signaling in the kidney.
Hypertension
17:
117-130,
1991[Abstract].
134.
Stutts, MJ,
Chinet TC,
Mason SJ,
Fullton JM,
Clarke LL,
and
Boucher RC.
Regulation of chloride channels in normal and cystic fibrosis airway epithelial cells by extracellular ATP.
Proc Natl Acad Sci USA
89:
1621-1625,
1992[Abstract].
135.
Stutts, MJ,
Fitz JG,
Paradiso AM,
and
Boucher RC.
Multiple modes of regulation of airway epithelial chloride secretion by extracellular ATP.
Am J Physiol Cell Physiol
267:
C1442-C1451,
1994
136.
Sugita, M,
Yue Y,
and
Foskett JK.
CFTR chloride channels and CFTR-associated ATP channels: distinct pores regulated by common gates.
EMBO J
17:
898-908,
1998
137.
Suko, Y,
Kawahara K,
Fukuda Y,
and
Masuda Y.
Nuclear and cytosolic calcium signaling induced by extracellular ATP in rat kidney inner medullary collecting duct cells.
Biochem Biophys Res Commun
234:
224-229,
1997[ISI][Medline].
138.
Superdock, KR,
Snyder DJ,
and
Breyer MD.
ATP-inhibitable Cl channel in apical membranes of cultured rabbit cortical collecting duct cells.
Am J Physiol Cell Physiol
265:
C957-C965,
1993
139.
Surprenant, A,
Rassendren F,
Kawashima E,
North RA,
and
Buell G.
The cytolytic P2Z receptor for extracellular ATP identified as a P2X receptor (P2X7).
Science
272:
735-738,
1996[Abstract].
140.
Takeda, M,
Kobayashi M,
and
Endou H.
Establishment of a mouse clonal early proximal tubule cell line and outer medullary collecting duct cells expressing P2 purinoceptors.
Biochem Mol Biol Int
44:
657-664,
1998[ISI][Medline].
141.
Tarasiuk, A,
Bar-Shimon M,
Gheber L,
Korngreen A,
Grossman Y,
and
Priel Z.
Extracellular ATP induces hyperpolarization and motility stimulation of ciliary cells.
Biophys J
68:
1163-1169,
1995[Abstract].
142.
Taylor, AL,
Kudlow BA,
Marrs KL,
Gruenert DC,
Guggino WB,
and
Schwiebert EM.
Bioluminescence detection of ATP release mechanisms in epithelia.
Am J Physiol Cell Physiol
275:
C1391-C1406,
1998
143.
Taylor, AL,
Schwiebert LM,
Smith JJ,
King C,
Jones JR,
Sorscher EJ,
and
Schwiebert EM.
Epithelial P2X purinergic receptor channel expression and function.
J Clin Invest
104:
875-84,
1999
144.
Thinnes, FP,
Hellmann KP,
Hellmann T,
Merker R,
Brockhaus-Pruchniewicz U,
Schwarzer C,
Walter G,
Gotz H,
and
Hilschmann N.
Studies of human porin XXII: cell membrane integrated human porin channels are involved in regulatory volume decrease (RVD) of HeLa cells.
Mol Genet Metab
69:
331-337,
2000[ISI][Medline].
145.
Torres, GE,
Egan TM,
and
Voigt MM.
Hetero-oligomeric assembly of P2X receptor subunits. Specificities exist with regard to possible partners.
J Biol Chem
274:
6653-6659,
1999
146.
Urano, T,
Nishimori H,
Han H,
Furuhata T,
Kimura Y,
Nakamura Y,
and
Tokino T.
Cloning of P2XM, a novel human P2X receptor gene regulated by p53.
Cancer Res
57:
3281-3287,
1997[Abstract].
147.
Van Calker, D,
Muller M,
and
Hamprecht B.
Adenosine regulates via two different types of receptors, the accumulations of cyclic AMP in cultured brain cells.
J Neurochem
33:
999-1005,
1979[ISI][Medline].
148.
Wang, DJ,
Huang NN,
and
Heppel LA.
Extracellular ATP and ADP stimulate proliferation of porcine aortic smooth muscle cells.
J Cell Physiol
153:
221-33,
1992[ISI][Medline].
149.
Wang, Y,
Roman RM,
Lidofsky SD,
and
Fitz JG.
Autocrine signaling through ATP release represents a novel mechanism for cell volume regulation.
Proc Natl Acad Sci USA
93:
12020-12025,
1996
150.
Wilson, PD.
Epithelial cell polarity and disease.
Am J Physiol Renal Physiol
272:
F434-F442,
1997
151.
Wilson, PD,
Hovater JS,
Casey CC,
Fortenberry JA,
and
Schwiebert EM.
ATP release mechanisms in primary cultures of epithelia derived from the cysts of polycystic kidneys.
J Am Soc Nephrol
10:
218-229,
1999
152.
Woda, CB,
Leit M,
and
Satlin LM.
Apical purinergic receptors in the rabbit cortical collecting duct (CCD) (Abstract).
J Am Soc Nephrol
11:
54A,
2000.
153.
Wong, LB,
and
Yeates DB.
Luminal purinergic regulatory mechanisms of tracheal ciliary beat frequency.
Am J Respir Cell Mol Biol
7:
447-454,
1992[ISI][Medline].
154.
Yamada, H,
Seki G,
Taniguchi S,
Uwatoko S,
Suzuki K,
and
Kurokawa K.
Mechanism of calcium increase by extracellular ATP in isolated rabbit proximal tubules.
Am J Physiol Cell Physiol
270:
C1096-C1104,
1996
155.
Zambon, AC,
Hughes RJ,
Meszaros JG,
Wu JJ,
Torres B,
Brunton LL,
and
Insel PA.
P2Y2 receptor of MDCK cells: Cloning, expression, and cell-specific signaling.
Am J Physiol Renal Physiol
279:
F1045-F1052,
2000
156.
Zhang, Y,
and
Engelhardt JF.
Airway surface liquid volume and chloride content in cystic fibrosis and normal bronchial xenografts.
Am J Physiol Cell Physiol
276:
C469-C476,
1999
157.
Zhou, QY,
Li C,
Olah ME,
Johsnson RA,
Stiles GL,
and
Civelli O.
Molecular cloning and characterization of an adenosine receptor: the A3 adenosine receptor.
Proc Natl Acad Sci USA
89:
7432-7436,
1992[Abstract].