Pore formation is not associated with macroscopic redistribution of P2X7 receptors

Megan L. Smart1,2, Rekha G. Panchal1,2, David N. Bowser1,2, David A. Williams2, and Steven Petrou1

1 The Laboratory of Biophysics and Molecular Physiology and 2 The Laboratory of Confocal and Fluorescence Imaging, Department of Physiology, The University of Melbourne, Victoria 3010, Australia


    ABSTRACT
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

The present study examines whether changes in P2X7 purinergic receptor density precede formation of the cytolytic pore characteristic of this receptor. We fused P2X7 receptors with enhanced green fluorescent protein (EGFP) at the amino or carboxy termini (EGFP-P2X7 and P2X7-EGFP). Electrophysiological characterization in Xenopus oocytes revealed wild-type responses to ATP for GFP-tagged receptors. However, differences in sensitivity to ATP were apparent with the P2X7-EGFP receptor displaying a threefold reduction in ATP sensitivity compared with control. Ethidium ion uptake was used to measure cytolytic pore formation. Comparison of tagged receptors with wild type in HEK-293 and COS-7 cells showed there was no significant difference in ethidium ion uptake, suggesting that fusions with EGFP did not interfere with cytolytic pore formation. Confocal microscopy confirmed that tagged receptors localized to the plasmalemma. Simultaneous monitoring of EGFP and ethidium ion fluorescence revealed that changes in receptor distribution do not precede pore formation. We conclude that it is unlikely that large scale changes in P2X7 receptor density precede pore formation.

pore-forming protein; cytolytic; HEK-293; P2Z; enhanced green fluorescent protein; fusion protein


    INTRODUCTION
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

THE P2X PURINERGIC RECEPTOR family contains seven members (P2X1 through P2X7). All members possess two transmembrane domains connected by a large extracellular loop and form ATP-gated ion channels (see Ref. 26 for review). The P2X7 receptor, first cloned from the rat brain (32), encodes a 595-amino acid protein that is structurally divergent from the other P2X subtypes, because it has a long (~200 amino acids) intracellular carboxy tail. This receptor is predominantly expressed in immune cells including macrophages, lymphocytes, and mast cells (5), although its expression is not restricted to cells of hemopoietic origin.

The activated P2X7 receptor has been shown to induce cell death either by an apoptotic (20) or necrotic pathway (34). It is this death-related function that implicates P2X7 receptor involvement in a number of pathological states including cancer (8), neurodegenerative diseases (12), and ischemic penumbras (25). Curiously, a benefit of this P2X7-induced cell death is the destruction of Mycobacterium tuberculosis residing within infected macrophages (7, 19).

A unique property of the P2X7 receptor is that it exhibits two agonist-activated conductance states. Upon activation with ATP or the potent agonist benzoylbenzoyl-ATP (BzATP), the receptor becomes permeant to small cations such as Na+, K+, and Ca2+ (34). Repeated or prolonged application of either agonist causes dilation of the pore, such that it is permeable to larger cations like ethidium ion or N-methyl-D-glucamine (11, 13, 27, 35). The mechanism underlying this pore dilation is unknown. Studies by Cockcroft and Gomperts (4) suggested that this dynamic selectivity is the result of successive oligomerization of monomeric subunits, resulting in a dilated pore permeable to larger cations. This is somewhat analogous to the dynamic oligomerization seen in a number of toxins (24, 31). An alternative mechanism is that a simple conformational change in the channel's selectivity filter is responsible for pore dilation or dynamic selectivity (14).

The capability of the P2X7 receptor to form pores (in endogenous cells) has been correlated with its relative surface expression, such that pore formation may require a threshold number of agonist-bound P2X7 receptor subunits (9). If this is indeed the case, then an increase in receptor density may encourage oligomerization events and subsequent pore formation.

Fluorescent tags such as green fluorescent proteins (GFPs; or variants of GFP) or fluorescently labeled antibodies have been used to examine receptor localization and receptor interactions with other proteins. Macroscopic changes such as alterations in receptor density, including receptor translocation to the cell membrane or real time receptor clustering, have been studied with tagged biomolecules (3, 36). In particular, GFP has been tagged to the carboxy termini of both the P2X1 and P2X2 receptors (6, 15) and expressed in neuronal cells. Interestingly, the P2X1-GFP construct was found to form "synaptic-size" clusters (1 µm) in the plasma membrane that were internalized on agonist addition (6). Examination of the P2X2-GFP chimera revealed that agonist addition induced varicose "hot spot" formation (1-2 µm) but not receptor internalization (15).

In the present study a similar approach was exploited to determine whether real time macroscopic clustering coincides or precedes pore formation of the rat P2X7 receptor. We define these clustering events as the appearance of localized areas of bright GFP fluorescence (hot spots) present on the plasmalemma that range from 1 to 2 µm in size. The formation of clusters may induce oligomerization of the P2X7 receptor, by increasing the probability of subunits to interact with one another because of an increase in receptor density. For this purpose, GFP-tagged P2X7 receptors were generated and their functional properties characterized. Pore formation was determined by ethidium ion uptake in both HEK-293 and COS-7 cells, and channel conductance was measured in Xenopus oocytes with the two-electrode voltage-clamp method. We examined whether GFP-tagged P2X7 receptors formed hot spots in HEK-293 and COS-7 cells, by performing a three-dimensional analysis of images taken with a confocal laser scanning microscope.


    EXPERIMENTAL PROCEDURES
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

Materials. pEGFP-N1 and pEGFP-C1 vectors were purchased from Clontech (California) and the pCI vector from Promega (Sydney, Australia). The Ambion T7 mMessage mMachine mRNA kit was sourced from Gene Works (Adelaide, Australia), and the Chroma Spin-100 columns were from Progen Industries (Brisbane, Australia). DMEM, penicillin-streptomycin, T4 DNA ligase, trypsin-EDTA, and ethidium bromide were purchased from Life Technologies (Melbourne, Australia). ATP, BzATP, poly-L-lysine, and trypan blue were all purchased from Sigma-Aldrich (Melbourne, Australia). Fetal bovine serum (FBS) and the Effectene transfection reagent kit were purchased from Commonwealth Serum Laboratories (Melbourne, Australia) and Qiagen (Melbourne, Australia), respectively.

Molecular biology. Enhanced GFP (EGFP)-tagged P2X7 receptors were created by subcloning the rat P2X7 receptor (rP2X7R) into the multiple cloning site of the pEGFP-N1 or the pEGFP-C1 vector. To generate the P2X7-EGFP construct, the rP2X7R was PCR amplified to introduce XhoI and KpnI restriction sites at the 5' and 3' ends, respectively, and to remove the stop codon. EGFP-P2X7 fusion was created by PCR amplification of rP2X7R to introduce the XhoI restriction site and delete the ATG start codon at the 5' end, with a HindIII site at the 3' end of the rP2X7R. The rat N-methyl-D-aspartate (NMDA)-1A-EGFP construct was created using PCR amplification, with a HindIII site at the 5' end and a BamHI site at the 3' end. All PCR products and the vectors were digested with appropriate restriction enzymes and ligated overnight using T4 DNA ligase. Ligated products were transformed into competent DH5alpha cells, and individual colonies were screened using standard restriction analysis procedures.

For electrophysiological studies, the P2X7-EGFP and EGFP-P2X7 constructs were subcloned into pCI and pcDNA3.1(+) vectors, respectively, and were linearized with the appropriate restriction enzyme in preparation for transcription runoff. In vitro transcription was performed using Ambion's T7 mMessage mMachine mRNA kit per the manufacturer's instructions. Full-length, capped cRNAs were purified using diethyl pyrocarbonate (DEPC) water-equilibrated Chroma Spin-100 columns. cRNA samples were electrophoresed on a denaturing agarose gel to determine the quantity, quality, and size of the transcript.

Cell culture. HEK-293 and COS-7 cells were grown at 37°C with 5% CO2 in DMEM, supplemented with 10% FBS and penicillin-streptomycin (100 U/ml). For imaging studies, the cells were trypsinized with trypsin-EDTA and seeded onto poly-L-lysine-treated coverslips. After 24 h, cells were transfected with P2X7-EGFP, EGFP-P2X7, or wild-type rP2X7R with Effectene reagent. To reduce cytotoxicity, HEK-293 cells were transfected for 5 h, washed, and then further incubated in complete DMEM medium. COS-7 cells were transfected overnight to increase transfection efficiency. Transfection efficiency was ~65-75% in HEK-293 cells but only 40% in COS-7 cells.

Confocal microscopy: receptor localization and pore formation. Transfected cells were viewed on a Bio-Rad MRC-1024 confocal microscope using the 488-nm line of the Argon laser. Fluorescent signals from EGFP and ethidium ion were viewed simultaneously in separate detector channels and sampled every 150 s. Time course data were collected using the Laser Sharp "Timecourse" software module (version 3.1). Recordings were made in a low divalent HEPES buffer (in mM: 147 NaCl, 2 KCl, 10 HEPES, 10 glucose, and 0.1 CaCl2, pH 7.4). ATP and BzATP were prepared in HEPES buffer (low divalent) containing ethidium bromide (10 µg/ml). Pore formation experiments were performed 48 h after transfection when plasmalemma localization in both HEK-293 and COS-7 cells was most evident. Ethidium ion uptake experiments were carried out at either room temperature (~24°C) or at 37°C. Experiments conducted at 37°C were performed using a water-jacketed immersion lens, and solutions were heated to the required temperature before experimentation. Ethidium ion fluorescence was measured in the nucleus of individual cells 10 min after agonist application. For each construct (at each agonist dose) four separate coverslips were analyzed, and on average 10 cells in a field were measured.

Three-dimensional images of transfected cells were obtained with an Olympus Fluoview laser scanning confocal microscope and software (FV500). Cells were imaged (Olympus ×40 water immersion lens, numerical aperture 0.9) for green fluorescence by excitation with the 488-nm line from an argon laser. A series of 30-40 images (Z-spacing 0.6 µm) was collected from a single cell, before and at specified time points (detailed in text) after agonist application. The two-dimensional tomographic images were reconstructed using the three-dimensional analysis module within the Fluoview software. Separate analysis of fluorescent beads (Molecular Probes, Eugene, Oregon) indicated that the optical system would be expected to resolve clusters >0.6 µm in diameter.

Trypan blue exclusion experiments. HEK-293 cells transfected with wild-type rP2X7R, P2X7-EGFP, or EGFP-P2X7 were tested for trypan blue exclusion on addition of BzATP (10 and 100 µM) at room temperature (~24°C). Uptake of ethidium ion and the visualization of trypan blue (0.4%) were monitored concurrently over a period of 2 h, using an epifluorescence microscope (Nikon Diaphot 300).

Expression and channel properties of EGFP-tagged rP2X7Rs. Oocytes from adult female Xenopus laevis (Disa Exporters, Somerset West Cape, South Africa) were surgically removed and prepared as outlined previously (29). Stage 5 or 6 oocytes were injected with cRNA encoding wild-type rP2X7R, P2X7-EGFP, or EGFP-P2X7 and were stored at 18°C for 48 h before experimentation. For two electrode voltage-clamp recordings, oocytes were impaled with glass electrodes containing 3 M KCl and held at a command potential of -70 mV using an Axoclamp 2B amplifier. Oocytes were perfused (at 2 ml/min) with a low divalent ND96 solution (in mM: 96 NaCl, 2 KCl, 0.1 BaCl2, and 5 HEPES, pH 7.5) using a gravity-fed manifold perfusion system. Different concentrations of ATP (0.33, 1, 3.3, 10, 33, 100, or 333 µM, or 1 mM) were applied to each oocyte until a peak current was achieved. Additional doses of ATP were applied after a 20-min washout period. EC50 values were determined from the responses of each oocyte.

Data analysis. Electrophysiological analysis of two-electrode voltage-clamp results was performed using CLAMPFIT (Axon Instruments) or ORIGIN 6.0 (Microal Software). Data in the text and figures are shown as means ± SE, and statistical analysis was performed using a Student's t-test. Data from ethidium ion uptake experiments were analyzed as described in the text and are presented as means ± SE. Statistical analysis was performed using a one-way ANOVA (Tukey's pairwise comparison).


    RESULTS
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

The P2X7-EGFP and EGFP-P2X7 constructs were designed to examine whether pore formation precedes or coincides with a change in receptor distribution, typically referred to as formation of receptor hot spots (15), which would be detectable with standard confocal microscopy. Before these constructs could be used for this purpose, it was important to characterize both the channel and pore-forming properties of the GFP-tagged receptors and compare them with properties of the wild-type rP2X7R. Channel properties were examined by two-electrode voltage-clamp analysis of Xenopus oocytes, and pore formation was studied by ethidium ion uptake in transfected HEK-293 or COS-7 cells.

Expression and channel properties of P2X7-EGFP, EGFP-P2X7, and wild-type rP2X7Rs. Application of 1 mM ATP to oocytes injected with either P2X7-EGFP, EGFP-P2X7, or wild-type rP2X7R cRNA induced a rapid inward current that recovered over a 5- to 6-min period, as illustrated in Fig. 1. To test the ATP sensitivity of fusion proteins, solutions with ATP in the concentration range from 0.33 µM to 1 mM were applied, and maximum peak currents were recorded and normalized to generate ATP dose-response curves. These curves revealed a reduced ATP sensitivity of the P2X7-EGFP receptor, compared with both EGFP-P2X7 and wild-type rP2X7R (Fig. 2). Altered receptor characteristics of P2X7-EGFP were further reflected in differences in the EC50 and the Hill coefficient. The EC50 value of wild-type rP2X7R was 4.9 ± 0.5 µM, compared with 6.0 ± 0.6 µM for EGFP-P2X7 and 25.5 ± 6.4 µM for P2X7-EGFP. The Hill coefficient calculated for P2X7-EGFP was 1.5 ± 0.2, which was significantly lower than that calculated for both wild-type rP2X7R (2.2 ± 0.2) and EGFP-P2X7 (2.5 ± 0.5). Interestingly, there was no significant difference in the maximum peak current between wild-type rP2X7R (2.84 ± 0.35 µA), P2X7-EGFP (2.76 ± 0.50 µA), and EGFP-P2X7 (2.72 ± 0.26 µA). Oocytes injected with P2X7-EGFP or EGFP-P2X7 cRNA exhibited green fluorescence when examined by confocal fluorescence microscopy. The application of BzATP to injected oocytes induced inward currents that generally took >30 min to recover (with low divalent conditions); hence, full BzATP dose-response curves were difficult to obtain, and ATP was used (29).


View larger version (6K):
[in this window]
[in a new window]
 
Fig. 1.   Comparison of ATP-induced current recordings of wild-type rat purinergic P2X7 receptor (rP2X7R) and enhanced green fluorescent protein (EGFP)-tagged P2X7 receptors expressed in Xenopus oocytes. Application of 1 mM ATP induced a rapid inward current that recovered after 5-6 min. The solid bar above each trace represents the duration of ATP application.



View larger version (13K):
[in this window]
[in a new window]
 
Fig. 2.   ATP dose-response curves for wild-type rP2X7R (n = 4), EGFP-P2X7 (n = 6), and P2X7-EGFP (n = 7) expressed in Xenopus oocytes. Oocytes injected with EGFP-P2X7 cRNA responded to ATP in the same manner as oocytes injected with wild-type rP2X7R cRNA. Oocytes injected with P2X7-EGFP cRNA displayed a decreased sensitivity to ATP, as demonstrated by the rightward shift, and a significantly different EC50 value of 25 µM (wild-type rP2X7R EC50 = 5 µM ) and Hill coefficient of 1.5 (wild-type rP2X7R, 2.2). Results are expressed as percentage of the maximal currents (Imax). Holding potential was -70 mV. I, current.

Oocytes injected with EGFP cRNA alone or with DEPC-treated water were tested as controls. Application of 1 mM ATP to either EGFP-injected oocytes (n = 7) or sham-injected oocytes (n = 5) elicited no response.

Ethidium ion uptake in HEK-293 or COS-7 cells. Pore formation was examined by a standard agonist-activated ethidium ion uptake procedure (13). HEK-293 cells transfected with either P2X7-EGFP, EGFP-P2X7, or wild-type rP2X7R accumulated ethidium ion on agonist application (10 and 100 µM BzATP; Fig. 3A). Ethidium ion fluorescence was augmented at 37°C, particularly in HEK-293 cells transfected with P2X7-EGFP. Simultaneous measurement of EGFP and ethidium ion fluorescence revealed that cells exhibiting EGFP fluorescence were also positive for ethidium ion fluorescence after prolonged application of BzATP indicative of pore formation (Fig. 3B), and within 30 min all EGFP-positive cells displayed an intense ethidium ion fluorescence in the nucleus. The bright ethidium ion fluorescence was not the result of membrane damage, because under the same conditions cells excluded trypan blue.


View larger version (30K):
[in this window]
[in a new window]
 
Fig. 3.   Fusion proteins expressed transiently in HEK-293 cells accumulate ethidium ion readily in response to benzoylbenzoyl-ATP (BzATP). A: both the fusion constructs and wild-type rP2X7R efficiently took up ethidium ion on BzATP application (10 or 100 µM). Increased uptake was observed at higher temperatures. Only the P2X7-EGFP receptor showed a significant increase in ethidium ion uptake at 37°C compared with room temperature (~24°C) at 10 µM BzATP concentration. *P < 0.05 (one-way ANOVA) significant difference of ethidium ion fluorescence between room temperature and 37°C. Values in parentheses indicate numbers of cells. B: simultaneous measurements of EGFP and ethidium ion (Ethidium+) fluorescence in HEK-293 cells expressing the P2X7-EGFP receptor. Top: in the absence of BzATP, EGFP fluorescence is readily observed and ethidium ion fluorescence is absent. Bottom: prolonged agonist activation (10 min, 10 µM BzATP) showed that cells expressing EGFP were also positive for ethidium ion fluorescence. Bar, 10 µm (all panels).

COS-7 cells expressing the fusion proteins or wild-type rP2X7R did not accumulate ethidium ion at room temperature (~24°C) with 10 µM BzATP, but significant uptake was observed at a higher concentration (100 µM) of BzATP (Fig. 4A). A shift in temperature to 37°C augmented ethidium ion uptake in COS-7 cells at both 10 and 100 µM concentrations of BzATP. The amount of ethidium ion uptake in COS-7 cells was always significantly less than that in HEK-293, as illustrated in Fig. 4B.


View larger version (24K):
[in this window]
[in a new window]
 
Fig. 4.   GFP-tagged receptors expressed transiently in COS-7 cells do not form pores readily in response to BzATP at room temperature. A: lower concentrations of BzATP (10 µM) produced ethidium ion uptake levels that were barely discernable above control background levels. Higher concentrations of BzATP (100 µM) or an increase in temperature (37°C) was required to obtain efficient ethidium ion uptake. Ethidium ion fluorescence in COS-7 cells was always significantly less intense than for equivalent measurements made in HEK-293 cells (compare with Fig. 3). *P < 0.05 (one-way ANOVA) significant difference of ethidium ion fluorescence between room temperature and 37°C. Values in parentheses indicate numbers of cells. B: simultaneous measurements of EGFP and ethidium ion fluorescence in COS-7 cells expressing the P2X7-EGFP receptor. Top: in the absence of BzATP, EGFP-positive cells excluded ethidium ion. Bottom: prolonged exposure (10 min) to a high concentration of BzATP (100 µM) was required for visualization of ethidium ion uptake, even though marked cell blebbing was observed. Bar = 10 µm (all panels).

Do receptor hot spots or increases in receptor density correlate with pore formation? Cellular localization of tagged rP2X7Rs was examined with confocal microscopy. Twenty-four hours after transfection of HEK-293 cells, the distribution of EGFP fluorescence was similar to that of the endoplasmic reticulum revealed by Di-O6 staining (data not shown). After 48 h, the fluorescence of both amino- and carboxy-terminal-tagged rP2X7Rs was apparent in the plasmalemma with some residual cytosolic signal (Fig. 5, A and B). In COS-7 cells, P2X7-EGFP was effectively targeted to the plasmalemma (Fig. 5C), whereas EGFP-P2X7 showed a more diffuse cytosolic localization (Fig. 5D). A number of multinucleated cells were present in transfected HEK-293 cells (expressing tagged or wild-type rP2X7Rs), as previously reported (2), but this phenomenon was never observed in transfected COS-7 cells.


View larger version (148K):
[in this window]
[in a new window]
 
Fig. 5.   Membrane localization of P2X7 fusion proteins in HEK-293 cells and COS-7 cells. HEK-293 cells transfected with P2X7-EGFP or EGFP-P2X7 receptors showed plasmalemma distribution with some diffused EGFP fluorescence in the cytosol (A and B). In COS-7 cells, P2X7-EGFP fluorescence was very distinct and localized to the plasmalemma (C). A more diffuse cytosolic fluorescence distribution was observed with EGFP-P2X7 (D). Bar = 10 µm.

The intensity of GFP fluorescence appeared to be increased at regions where HEK-293 cells made contact with each other (Fig. 6A). Analysis of three-dimensional reconstructions of image stacks for cells in contact could not unequivocally determine whether the increased intensity was due to a greater density of P2X7-EGFP receptors in the contact region or whether it reflected the complex contact morphology and its sometimes planar contribution to two-dimensional images. However, it was clear that receptor clusters were not evident in these regions (Fig. 6B).


View larger version (49K):
[in this window]
[in a new window]
 
Fig. 6.   GFP-fused P2X7 receptors do not form clusters in regions of cell-to-cell contact. Some HEK-293 cells transfected with P2X7-EGFP (A) reveal an apparent increase in GFP fluorescence intensity at the site of cell-to-cell contact when the three-dimensional image stacks are rotated (B). Arrow, site of cell-to-cell contact. Hot spot or cluster formation was not observed in these regions. Bar = 10 µm. The localized fluorescent signal seen within the cells was not membrane delimited; this signal was from endoplasmic reticulum or Golgi-bound receptors.

To determine whether receptor hot spot formation coincided with or preceded pore formation, EGFP fluorescence was monitored in HEK-293 cells transfected with either the P2X7-EGFP or EGFP-P2X7 tagged receptor. For each cell, a Z-stack was collected before the addition of BzATP (10 µM) and then at 5, 10, and 30 min after agonist application. Receptor hot spots were not evident before or after agonist addition in HEK-293 cells transfected with the P2X7-EGFP (Fig. 7A) or EGFP-P2X7 (Fig. 7B) receptor (6 individual coverslips examined for each construct; n = 6 cells). Images of the NMDA-1A-EGFP receptor distribution in HEK-293 cells (Fig. 7C) were used as a positive control and provided clear evidence for clustering under basal conditions (22). Receptor clusters of 1-2 µm in diameter have been readily resolved with both P2X1 and P2X2-GFP tagged receptors (6, 15), suggesting that, if such events did occur (as observed in our hands with the NMDA-1A-EGFP receptor in HEK-293 cells) with the GFP-fused P2X7 receptors, they should have been readily observed in the basal or activated state.


View larger version (41K):
[in this window]
[in a new window]
 
Fig. 7.   Addition of BzATP to HEK-293 cells transfected with P2X7-EGFP and EGFP-P2X7 receptors did not induce the formation of hot spots. Three-dimensional image stacks were obtained before the addition of BzATP (10 µM) and then 5, 10, and 30 min later. The two-dimensional image views of the Z-stack are presented for HEK-293 cells transfected with P2X7-EGFP (A) and EGFP-P2X7 (B). While gross morphological changes such as blebbing, microvesiculation, and ultimately cell death were observed, hot spots were not apparent before or after agonist addition. The N-methyl-D-aspartate (NMDA)-1A-EGFP receptor when expressed in HEK-293 cells (C) displayed clear evidence of receptor hot spots (~0.6 µm in size). These hot spots represent small clusters of the EGFP-tagged NMDA receptor. Bar = 10 µm.

Although hot spot formation does not occur in HEK-293 cells transfected with GFP-tagged rP2X7Rs, a series of morphological changes was observed in stimulated cells. Such responses were cytoplasmic microvesiculation (evident in cells with a significant amount of EGFP within the endoplasmic reticulum) and membrane blebbing. This dramatic morphological change was not as obvious in COS-7 cells, and these cells did not display any GFP-tagged P2X7 receptor clusters (data not shown).


    DISCUSSION
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

The mechanism by which agonists such as ATP induce P2X7 receptors to change ion size selectivity is still not known. A number of intra- and intermolecular models have been proposed, which describe the process that accounts for the change in ion selectivity of this receptor. Intramolecular models suggest that subtle structural changes, possibly in the selectivity filter itself, contribute to the dynamic change from channel to pore state (14). Intermolecular models predict that pore formation occurs by dilation of the channel, wherein successive addition of extra subunits produces an oligomeric pore capable of conducting large size (~900 Da) cations (4). This latter possibility was the focus of the present studies. If a real time oligomerization is responsible for pore dilation, then real time subunit redistribution or clustering may be visualized by imaging EGFP-tagged rP2X7Rs. Because it was not known whether EGFP tagged to either the amino or carboxy terminus would affect the functional properties of the P2X7 receptor, we examined the functional consequences of fusing EGFP to either end of the rP2X7R.

Two-electrode voltage-clamp recordings revealed that ATP-induced currents of the fusion constructs were similar to those of wild-type rP2X7R and to previously published work (29). Wild-type rP2X7R displayed higher sensitivity to ATP when expressed in oocytes (29) than in HEK-293 cells (32), although activation curves of both systems were best described by Hill coefficients of ~2 (30). P2X7-EGFP fusion was shown to have a decreased ATP sensitivity compared with both wild-type or EGFP-P2X7 receptors. The decreased ATP sensitivity of P2X7-EGFP indicates that EGFP tagged to the carboxy terminus disrupts either the binding of ATP to the receptor or alters transduction subsequent to the binding event. A disruption of ATP binding is unlikely, since ATP binding occurs within the extracellular loop (10) and both the carboxy and amino terminals are cytosolic. This result contributes to the growing evidence of the involvement of the carboxy terminal tail in regulating the function of the P2X7 receptor.

The absolute requirement of the full-length carboxy-terminal tail for P2X7 pore formation has been previously established (32). Until recently, little was known about the involvement of the carboxy-terminal tail with respect to the channel response. Klapperstuck and colleagues (16) observed in Xenopus oocytes (expressing the human P2X7 receptor) that the carboxy-terminal tail is involved in the regulation of ATP-induced (channel) currents, as the truncated receptor (human P2X7 1-418) yielded an altered response that differed from that of the wild-type receptor. Unlike other members of the P2X family (P2X1-4), where a string of charged residues within the carboxy-terminal tail has been found to alter channel kinetics (17), the carboxy-terminal tail of the P2X7 receptor appears to be more complex, with the function of crucial regions largely unresolved.

Previous studies have also shown that ion channels tagged with GFP exhibit an altered response. In particular, Khakh and colleagues (15) revealed a rightward shift in the ATP dose-response curve of the P2X2-GFP construct when expressed in Xenopus oocytes, although there was no change in the Hill coefficient. GFP fused to the amino terminus altered the kinetics of K+ channels, although GFP addition to the carboxy terminus was not examined (18, 23). The present study highlights the importance of functional testing of both amino- and carboxy-terminal fusion constructs before other applications.

Pore formation was examined by ethidium dye uptake measurements in the continued presence of BzATP. Both amino- and carboxy-terminal tagged rP2X7Rs were permeable to ethidium ion, with the amount of ethidium ion uptake (10 min post-BzATP addition) comparable to that of wild-type rP2X7R. The observed difference in ethidium ion uptake between HEK-293 and COS-7 cells suggests that rP2X7R function may be regulated by local factors such as lipid environment or involvement of a cell-specific protein partner. In COS-7 cells ethidium ion uptake was augmented by increasing the temperature from ~24°C to 37°C, in agreement with earlier work by Cario-Toumaniantz and colleagues (1).

The involvement of a protein partner has been previously postulated as an explanation for the variant ion size selection of the P2X7 receptor, between cells and species. As yet little is known about the protein-protein interactions of the P2X7 receptor. Postulated protein partners might play the role of pore-enabling or pore-inhibiting factor. The P2X7 receptor is known to form homooligomers, but heteroligomers with other members of the P2X family are not apparent (33). Interestingly, Parker (28) observed that cytoskeletal actin interacts with and regulates the kinetics of the P2X1 receptor. Likewise McKenzie and Surprenant (21) observed that F actin is involved in the blebbing induced by P2X7 activation. Further investigation into the regulation of pore formation by P2X7 protein partners or the interaction of the receptor with cytoskeletal proteins is warranted.

Confocal imaging studies showed plasmalemma localization of rP2X7R in HEK-293 and COS-7 cells. However, the location of the EGFP does affect the expression pattern. EGFP-P2X7 expressed in HEK-293 cells could be localized to the plasmalemma, although this was not the case when this construct was expressed in COS-7 cells. This difference is intriguing and may point to the involvement of protein partners that are required for trafficking and expressed at varying levels in different cell types.

An increase in membrane density of rP2X7R has been proposed to augment pore formation (9). It is peculiar that COS-7 cells expressing P2X7-EGFP exhibit intense plasmalemma fluorescence but only a mild pore-forming phenotype, suggesting that there is no simple relationship between receptor density and propensity to form pores.

Recent studies examining the localization of the P2X1-GFP fusion protein demonstrated synaptic size clusters in the plasma membrane of HEK-293 cells and dissociated neurons (6), although these receptor clusters were present before agonist application. Agonist-induced changes in GFP-tagged rat P2X2 receptors expressed in neuronal cells could be readily observed in dendrites, demonstrating that real time receptor redistribution is a physiologically significant response to receptor activation (15). In contrast, this study reveals that P2X7 pore formation does not involve either macroscopic receptor density changes or receptor clustering.


    NOTE ADDED IN PROOF

A proteomic analysis of P2X7 protein partners has been recently published: Kim M, Jiang LH, Wilson HL, North RA, and Surprenant A. Proteomic and functional evidence for a P2X7 receptor signalling complex. EMBO J 20: 6347-6358, 2001.


    ACKNOWLEDGEMENTS

We thank both reviewers for their helpful and insightful comments.


    FOOTNOTES

This project was supported by the Anti Cancer Council of Victoria (Australia) and the William Buckland Foundation (ANZ Trustees, Australia).

Address for reprint requests and other correspondence: S. Petrou, The Laboratory of Biophysics and Molecular Physiology, The Univ. of Melbourne, Victoria 3010, Australia (E-mail: spetrou{at}unimelb.edu.au).

The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

First published February 13, 2002;10.1152/ajpcell.00456.2001

Received 24 September 2001; accepted in final form 12 February 2002.


    REFERENCES
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

1.   Cario-Toumaniantz, C, Loirand G, Ladoux A, and Pacaud P. P2X7 receptor activation-induced contraction and lysis in human saphenous vein smooth muscle. Circ Res 83: 196-203, 1998[Abstract/Free Full Text].

2.   Chiozzi, P, Sanz JM, Ferrari D, Falzoni S, Aleotti A, Buell GN, Collo G, and Di Virgilio F. Spontaneous cell fusion in macrophage cultures expressing high levels of the P2Z/P2X7 receptor. J Cell Biol 138: 697-706, 1997[Free Full Text].

3.   Ciruela, F, Saura C, Canela EI, Mallol J, Lluis C, and Franco R. Ligand-induced phosphorylation, clustering, and desensitization of A1 adenosine receptors. Mol Pharmacol 52: 788-797, 1997[Abstract/Free Full Text].

4.   Cockcroft, S, and Gomperts BD. ATP induces nucleotide permeability in rat mast cells. Nature 279: 541-542, 1979[ISI][Medline].

5.   Collo, G, Neidhart S, Kawashima E, Kosco-Vilbois M, North RA, and Buell G. Tissue distribution of the P2X7 receptor. Neuropharmacology 36: 1277-1283, 1997[ISI][Medline].

6.   Dutton, JL, Poronnik P, Li GH, Holding CA, Worthington RA, Vandenberg RJ, Cook DI, Barden JA, and Bennett MR. P2X(1) receptor membrane redistribution and down-regulation visualized by using receptor-coupled green fluorescent protein chimeras. Neuropharmacology 39: 2054-2066, 2000[ISI][Medline].

7.   Fairbairn, IP, Stober CB, Kumararatne DS, and Lammas DA. ATP-mediated killing of intracellular mycobacteria by macrophages is a p2x(7)-dependent process inducing bacterial death by phagosome-lysosome fusion. J Immunol 167: 3300-3307, 2001[Abstract/Free Full Text].

8.   Gu, BJ, Zhang WY, Bendall LJ, Chessell IP, Buell GN, and Wiley JS. Expression of P2X(7) purinoceptors on human lymphocytes and monocytes: evidence for nonfunctional P2X(7) receptors. Am J Physiol Cell Physiol 279: C1189-C1197, 2000[Abstract/Free Full Text].

9.   Gudipaty, L, Humphreys BD, Buell G, and Dubyak GR. Regulation of P2X(7) nucleotide receptor function in human monocytes by extracellular ions and receptor density. Am J Physiol Cell Physiol 280: C943-C953, 2001[Abstract/Free Full Text].

10.   Hansen, MA, Barden JA, Balcar VJ, Keay KA, and Bennett MR. Structural motif and characteristics of the extracellular domain of P2X receptors. Biochem Biophys Res Commun 236: 670-675, 1997[ISI][Medline].

11.   Hibell, AD, Thompson KM, Simon J, Xing M, Humphrey PP, and Michel AD. Species- and agonist-dependent differences in the deactivation-kinetics of P2X7 receptors. Naunyn Schmiedebergs Arch Pharmacol 363: 639-648, 2001[ISI][Medline].

12.   Hide, I, Tanaka M, Inoue A, Nakajima K, Kohsaka S, Inoue K, and Nakata Y. Extracellular ATP triggers tumor necrosis factor-alpha release from rat microglia. J Neurochem 75: 965-972, 2000[ISI][Medline].

13.   Humphreys, BD, and Dubyak GR. Induction of the P2z/P2X7 nucleotide receptor and associated phospholipase D activity by lipopolysaccharide and IFN-gamma in the human THP-1 monocytic cell line. J Immunol 157: 5627-5637, 1996[Abstract].

14.   Khakh, BS, and Lester HA. Dynamic selectivity filters in ion channels. Neuron 23: 653-658, 1999[ISI][Medline].

15.   Khakh, BS, Smith WB, Chiu CS, Ju D, Davidson N, and Lester HA. Activation-dependent changes in receptor distribution and dendritic morphology in hippocampal neurons expressing P2X2-green fluorescent protein receptors. Proc Natl Acad Sci USA 98: 5288-5293, 2001[Abstract/Free Full Text].

16.   Klapperstuck, M, Buttner C, Schmalzing G, and Markwardt F. Functional evidence of distinct ATP activation sites at the human P2X7 receptor. J Physiol 534: 25-35, 2001[Abstract/Free Full Text].

17.   Koshimizu, T, Koshimizu M, and Stojilkovic SS. Contributions of the C-terminal domain to the control of P2X receptor desensitization. J Biol Chem 274: 37651-37657, 1999[Abstract/Free Full Text].

18.   Kupper, J. Functional expression of GFP-tagged Kv1.3 and Kv14 channels in HEK 293 cells. Eur J Neurosci 10: 3908-3912, 1998[ISI][Medline].

19.   Lammas, DA, Stober C, Harvey CJ, Kendrick N, Panchalingam S, and Kumararatne DS. ATP-induced killing of mycobacteria by human macrophages is mediated by purinergic P2Z(P2X7) receptors. Immunity 7: 433-444, 1997[ISI][Medline].

20.   Le Feuvre, RA, Brough D, Iwakura Y, Takeda K, and Rothwell NJ. Priming of macrophages with lipopolysaccharide potentiates P2X7-mediated cell death via a caspase-1-dependent mechanism, independently of cytokine production. J Biol Chem 277: 3210-3218, 2001.

21.   MacKenzie, A, and Surprenant A. A role for the actin cytoskeleton in P2X7 receptor-mediated changes in HEK-293 cell morphology. (Abstract). J Physiol 525: 74P, 2000.

22.   Matsuda, S, and Hirai H. The clustering of NMDA receptor NR1 subunit is regulated by the interaction between the C-terminal exon cassettes and the cytoskeleton. Neurosci Res 34: 157-163, 1999[ISI][Medline].

23.   Meyer, E, and Fromherz P. Ca2+ activation of hSlo K+ channel is suppressed by N-terminal GFP tag. Eur J Neurosci 11: 1105-1108, 1999[ISI][Medline].

24.   Mogridge, J, Mourez M, and Collier RJ. Involvement of domain 3 in oligomerization by the protective antigen moiety of anthrax toxin. J Bacteriol 183: 2111-2116, 2001[Abstract/Free Full Text].

25.   Morigiwa, K, Fukuda Y, and Yamashita M. Neurotransmitter ATP and cytokine release. Nippon Yakurigaku Zasshi 115: 185-192, 2000[Medline].

26.   North, RA, and Surprenant A. Pharmacology of cloned P2X receptors. Annu Rev Pharmacol Toxicol 40: 563-580, 2000[ISI][Medline].

27.   Nuttle, LC, and Dubyak GR. Differential activation of cation channels and non-selective pores by macrophage P2z purinergic receptors expressed in Xenopus oocytes. J Biol Chem 269: 13988-13996, 1994[Abstract/Free Full Text].

28.   Parker, KE. Modulation of ATP-gated non-selective cation channel (P2X1 receptor) activation and desensitization by the actin cytoskeleton. J Physiol 510: 19-25, 1998[Abstract/Free Full Text].

29.   Petrou, S, Ugur M, Drummond RM, Singer JJ, and Walsh JV. P2X7 purinoceptor expression in Xenopus oocytes is not sufficient to produce a pore-forming P2Z-like phenotype. FEBS Lett 411: 339-345, 1997[ISI][Medline].

30.   Pizzo, P, Zanovello P, Bronte V, and Di Virgilio F. Extracellular ATP causes lysis of mouse thymocytes and activates a plasma membrane ion channel. Biochem J 274: 139-144, 1991[ISI][Medline].

31.   Sharpe, JC, and London E. Diphtheria toxin forms pores of different sizes depending on its concentration in membranes: probable relationship to oligomerization. J Membr Biol 171: 209-221, 1999[ISI][Medline].

32.   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].

33.   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[Abstract/Free Full Text].

34.   Virginio, C, MacKenzie A, North RA, and Surprenant A. Kinetics of cell lysis, dye uptake and permeability changes in cells expressing the rat P2X7 receptor. J Physiol 519: 335-346, 1999[Abstract/Free Full Text].

35.   Wiley, JS, Gargett CE, Zhang W, Snook MB, and Jamieson GP. Partial agonists and antagonists reveal a second permeability state of human lymphocyte P2Z/P2X7 channel. Am J Physiol Cell Physiol 275: C1224-C1231, 1998[Abstract/Free Full Text].

36.   Wulfing, C, Sjaastad MD, and Davis MM. Visualizing the dynamics of T cell activation: intracellular adhesion molecule 1 migrates rapidly to the T cell/B cell interface and acts to sustain calcium levels. Proc Natl Acad Sci USA 95: 6302-6307, 1998[Abstract/Free Full Text].


Am J Physiol Cell Physiol 283(1):C77-C84
0363-6143/02 $5.00 Copyright © 2002 the American Physiological Society