AuPS Logo Contents
Previous Next PDF

Store-operated Ca2+ channels and microdomains of Ca2+ in liver cells

Greg J. Barritt,* Tom L. Litjens, Joel Castro,* Edoardo Aromataris and Grigori Y. Rychkov
*Department of Medical Biochemistry, School of Medicine, Flinders University, PO Box 2100, Adelaide, SA 5001, Australia and Department of Physiology, School of Molecular and Biomedical Science, University of Adelaide, Adelaide, SA 5005, Australia

Summary

1. Oscillatory increases in the cytoplasmic Ca2+ concentration ([Ca2+]cyt) play essential roles in the hormonal regulation of liver cells. Increases in [Ca2+]cyt require Ca2+ release from the endoplasmic reticulum (ER) and Ca2+ entry across the plasma membrane.

2. Store-operated Ca2+ channels (SOCs), activated by a decrease in Ca2+ in the ER lumen, are responsible for maintaining adequate ER Ca2+. Experiments employing patch clamp recording and the fluorescent Ca2+ reporter fura-2 indicate there is only one type of SOC in rat liver cells. These SOCs have a high selectivity for Ca2+ and properties essentially indistinguishable from those of Ca2+ release-activated Ca2+ (CRAC) channels.

3. While Orai1, a CRAC channel pore protein, and Stim1, a CRAC channel Ca2+ sensor, are components of the liver cell SOCs, the mechanism of activation of SOCs, and in particular the role of subregions of the ER, are not well understood.

4. Recent experiments have employed the transient receptor potential V1 (TRPV1) non-selective cation channel, ectopically expressed in liver cells, and a choleretic bile acid to deplete Ca2+ from different ER subregions. The results have provided evidence that only a small component of the ER is required for STIM1 re-distribution and the activation of SOCs.

5. It is concluded that different Ca2+ microdomains in the ER and cytoplasmic space are important in both the activation of SOCs and in the signalling actions of Ca2+ in liver cells. Future experiments will further investigate the nature of these microdomains.

Introduction

Hormone-induced increases in the concentration of Ca2+ in the cytoplasmic space ([Ca2+]cyt) play a central role in intracellular signalling in animal cells.1 In liver cells, one of the first types of animal cell in which oscillations in [Ca2+]cyt were observed,2 hormone-induced oscillations in [Ca2+]cyt regulate pathways of intermediary and xenobiotic metabolism, bile acid secretion, cell proliferation and apoptosis and necrosis.3 The maintenance of hormone-induced Ca2+ oscillations requires the constant replenishment of the endoplasmic reticulum (ER) Ca2+ stores by Ca2+ entering the cell through Ca2+ entry channels in the plasma membrane. While several types of Ca2+ entry channel might be involved in maintaining adequate Ca2+ in the ER, store-operated Ca2+ channels (SOCs) play a major role. The activation of SOCs is initiated by a decrease in Ca2+ in the ER. In hormonally-stimulated liver cells and in other animal cells, this ER Ca2+ decrease is mediated by inositol 1,4,5-trisphosphate receptors (IP3Rs) (reviewed by Parekh & Putney4). Experimentally ER Ca2+ release can be induced by inhibition of the (Ca2+ + Mg2+)ATPase in the ER (SERCA) with thapsigargin, 2,5-di-(tert)-butyl)-1,4-benzohydro-quinone (DBHQ) and other SERCA inhibitors. The most extensively characterised SOCs are the Ca2+ release-activated Ca2+ channels (CRAC) in lymphocytes and mast cells (reviewed by Parekh & Putney4). The aim of this short review is to summarize the properties of SOCs in liver cells, current knowledge of their molecular components, likely mechanisms of activation, and the roles of Ca2+ microdomains in the activation mechanism and in the regulation of Ca2+ entry.

Characteristics of store-operated Ca2+ channels in liver cells

Many studies have shown that a SERCA inhibitor or IP3 (introduced by micro injection or generated by addition of a hormone) will initiate the activation of Ca2+ entry to liver cells.5-16 Since SOCs have often been functionally defined as channels which are activated by treatment of cells with SERCA inhibitor or IP34 Ca2+ entry in response to these agents has often been attributed to SOCs. Moreover, it was suggested that more than one type of SOC may be expressed in hepatocytes and liver cell lines.8,10,13,17 However, in the majority of these studies the nature of the Ca2+ entry pathway involved was not clearly defined. In recent patch clamp experiments with rat liver cells only one type of SOC, a highly Ca2+-selective channel similar to CRAC channels, could be detected.18-20 It is possible that in some earlier studies IP3 and SERCA inhibitors may have initiated the activation of non-SOCs (reviewed by Barritt et al.3).

Ca2+-permeable channels activated by thapsigargin and IP3 in H4-IIE rat liver cells and rat hepatocytes have been characterised using patch clamp recording.18-20 These SOCs exhibit a high selectivity for Ca2+ compared with monovalent cations and exhibit properties similar, or identical, to those of the CRAC channels found in lymphocytes and mast cells.18-20 Time courses of activation, current amplitudes, dependence on the extracellular Ca2+ concentration ([Ca2+]ext), conductance for Ba2+ compared with Ca2+, and inhibition by La3+, Gd3+ and 2-aminoethyl diphenylborate (2-APB) are similar for liver cell SOCs and CRAC channels. Ca2+ entry, measured by whole cell patch clamp recording, through SOCs in rat hepatocytes can be activated by physiological concentrations of vasopressin and ATP.20 The permeability sequence for the movement of cations through liver cell SOCs is Ca2+ > Ba2+ > Sr2+ > Na+ > Cs+.18

Liver cell SOCs are partially blocked by Co2+ and Cd2+ and completely blocked by Zn2+, Gd3+ and La3+. The most potent blocking agents are Gd3+ and La3+ which give complete block at about 2 μM in the presence of 10 mM Ca2+ext.18,20 The concentration of 2-APB which gives half-maximal inhibition of Ca2+ entry is approximately 10 μM.21 Ca2+ entry through liver cell SOCs is also inhibited by SK&F 96365,15,22 arachidonic acid,23 the phospholipase C (PLC) inhibitor U73122,24,25 and by isotetrandrine and tetrandrine.23

There is some evidence to indicate that calmodulin is involved in the regulation of liver cell SOCs.5,22 The results of patch clamp studies with H4-IIE rat liver cells have provided evidence that the fast inactivation of the SOC Ca2+ current, ISOC, is a calmodulin- and Ca2+-dependent process, similar to the Ca2+-dependent fast inactivation of CRAC channels.19 Thus over-expression of either a calmodulin inhibitor peptide or a mutant form of calmodulin lacking functional EF hand domains reduced the fast component of liver cell ISOC inactivation. However, no effect of the calmodulin antagonists Mas-7 and calmidazolium was detected. It is possible that calmodulin is tethered to the rat liver SOC protein which shields it from the actions of calmodulin inhibitors.26

Molecular components of store-operated Ca2+ channels in liver cells

Experiments conducted during the past 3 years have shown that a member of the Orai (CRACM) family of proteins constitutes the pore of SOCs in mast cells, lymphocytes and in many other types of animal cells while STIM1 (stromal interaction factor 1) located in the ER constitutes the Ca2+ sensor. STIM1 is thought to detect the decrease in [Ca2+]er and convey this information to Orai at the plasma membrane leading to activation of the channel and Ca2+ entry. This involves the movement of some STIM to ER-plasma membrane junctions leading to an interaction between STIM and Orai.27-31 Further experiments are required to determine whether there is a direct interaction between STIM and Orai, or whether additional proteins are involved. The localisation of Orai and STIM and the Ca2+ entry channel may create domains of increased [Ca2+]cyt at specific locations under the plasma membrane.28

It is likely that STIM is required in the mechanism of activation of liver cell SOCs. Knockdown of STIM1 in H4-IIE liver cells using siRNA caused a substantial reduction in the amplitude of ISOC initiated by IP3 or thapsigargin.25 Treatment of H4-IIE cells with thapsigargin led to a redistribution of STIM1 to puncta, as assessed using cells transfected with GFP-STIM1 and by imaging endogenous STIM1 by immunofluorescence.32

The proposed mechanism of activation of liver cell SOCs involving the interaction of STIM1 with Orai1 at ER-plasma membrane junctions requires that such junctions are normally present in hepatocytes or are formed upon depletion of Ca2+ in the ER. Evidence for a close association of some ER with the plasma membrane in hepatocytes comes from previous subcellular fractionation experiments which generated highly purified plasma membrane fractions and provided evidence that specialised subregions of the ER are located close to the plasma membrane.33,34

In some other types of animal cells, TRP (transient receptor potential) proteins, including TRPC1, TRPC3, TRPC4, TRPV5 and/or TRPV6, are thought to constitute the pores of SOCs.4,35,36 Some of these TRP proteins are expressed in liver cells (reviewed by Barritt et al.3). In H4-IIE rat liver cells, ectopic expression of hTRPC1 or knockdown of endogenous TRPC1 proteins using siRNA did not substantially change thapsigargin-stimulated Ca2+ entry (assessed using a fluorescent Ca2+ sensor and patch clamp recording), indicating that it is unlikely that the TRPC1 peptide is a component of SOCs in rat liver cells.37,38 As described above, in patch clamp recording experiments only one type of SOC can be detected in rat liver cells and this has a high selectivity for Ca2+, comparable to that of CRAC channels in lymphocytes and mast cells. The Ca2+-permeable channels formed by TRPC1 polypeptides and by most other TRP polypeptides have a relatively low selectivity for Ca2+ compared with Na+.39,40 This suggests that it is unlikely that any of the known TRP polypeptides constitutes the Ca2+-selective SOCs found in rat hepatocytes and liver cells.

Trimeric GTP-binding protein Gi2, F-actin and phospholipase Cγ1 may play permissive roles in the activation of liver cell SOCs

While STIM1 and Orai1 proteins are likely to be the major proteins which constitute liver cell SOCs, several other proteins appear to be required. Knockdown of PLCγ1 in H4-IIE rat liver cells using siRNA was found to be associated with a substantial decrease in the amplitude of ISOC initiated by either IP3 or thapsigargin. No interaction between PLCγ1 and STIM1 was detected in immunoprecipitation experiments.25 It was concluded that PLC-γ1 is required to couple ER Ca2+ release to the activation of SOCs independently of any PLCγ1-mediated generation of IP3 and independently of a direct interaction between PLCγ1 and STIM1.

ADP-ribosylation of the trimeric GTP-binding protein, Gi2α, by treatment of livers with pertussis toxin, or the inhibition of Gi2α function using an inhibitory antibody or an inhibitory peptide, were each found to inhibit thapsigargin- and IP3-induced Ca2+ entry (measured using fura-2) to freshly-isolated rat hepatocytes.41-45 ADP-ribosylation of Gi2α was associated with inhibition of the formation of the band of cortical F-actin around the canaliculus in isolated hepatocyte doublets when spatial polarity was regained, and with some disruption of the ER.46 Moreover, studies with hepatocytes have shown that Gi2α interacts with F-actin.46 Disruption of F-actin with cytochalasin D, within a narrow concentration range, inhibited thapsigargin- and IP3-induced Ca2+ entry.47 Taken together, these results indicate that the normal functions of Gi2α and F-actin are required for the activation of hepatocyte SOCs. The results of other studies suggest that, in addition to Gi2, a monomeric G-protein, possible ARF-1, is also required for the activation of SOCs in hepatocytes.48

Since the interventions described above inhibited the activation of SOCs when this was initiated by thapsigargin as well as by IP3, it was concluded that the requirements for Gi2α and F-actin are downstream of the step in which Ca2+ is released from the ER. Thus, it was proposed that Gi2α is not involved in the formation of IP3, catalysed by PLCβ linked to a G protein-coupled receptor, but rather that the role of Gi2α in the activation of SOCs represents a “receptor independent” function of Gi2α (cf. the role of the Gi3 in vesicle trafficking and in other receptor-independent functions of G-proteins.49 Gi2 may function to maintain hepatocyte spatial polarity since it has been shown that trimeric G-proteins are involved in determining cell polarity in other cell types.50 PLCγ1, Gi2α, the monomeric G protein and F-actin may play “permissive” roles, such as maintenance of the integrity of the ER and the putative ER-plasma membrane junctions, in SOC activation in spatially polarised hepatocytes.

Roles of IP3 receptor subtypes and putative subregions of the endoplasmic reticulum in the activation of liver cell SOCs

Two questions concerning the roles of IP3R and the ER in the activation of SOCs in liver cells have been addressed. The first is whether a specific subtype of IP3R is required for SOC activation, and the second is whether all of the ER or only a sub-component of the ER is required for the activation of SOCs. Rat hepatocytes express type 1 (20%), type 2 (80%) and a small proportion (<1%) of type 3 inositol 1,4,5-trisphosphate receptors (IP3Rs).51-53 In hepatocytes, type 2 IP3Rs are expressed chiefly in the pericanalicular region and are responsible for the initiation of waves of increased [Ca2+]cyt originating from this region.51,53,54 When microinjected into freshly-isolated hepatocytes, a monoclonal anti-type 1 IP3R antibody, which in other studies was shown to inhibit Ca2+ release mediated by type 1 IP3R, was found to inhibit hormone- and thapsigargin-induced Ca2+ entry with little effect on the release of Ca2+ from intracellular stores.55 The microinjection of a relatively low concentration of adenophostin A, which has a high affinity for IP3Rs relative to that of IP3, induced near-maximal activation of Ca2+ entry with little detectable release of Ca2+ from intracellular stores.55 The results of experiments in which IP3 analogues selective for either type 1 or type 2 IP3R were microinjected to rat hepatocytes suggest that type 1 IP3R are preferentially involved in SOC activation.56

As mentioned above, type 2 IP3R are predominantly located in the ER near the bile canaliculus while type 1 IP3R are distributed throughout most regions of the ER with some type 1 IP3R concentrated in ER close to the plasma membrane in the sinusoidal and canalicular domains.51,54,56 The results of subcellular fractionation studies indicate that type 1 IP3R are found in regions of the ER very close to the plasma membrane, and are held in this location by F-actin.33,34,57 Taken together, the results obtained using these different experimental approaches suggest that a small subregion of the ER enriched in type 1 IP3Rs is required for SOC activation.

The question of whether the activation of liver cell SOCs requires the whole or only a small component of the ER has been further investigated using the non-selective cation channel TRPV1 and the choleretic bile acid taurodeoxycholic acid (TDCA) to release Ca2+ from different regions of the ER. It has previously been shown that TDCA activates SOCs in liver cells by releasing Ca2+ from the ER and causing a redistribution of STIM1.32 When ectopically expressed in H4-IIE rat liver cells, TRPV1 was found to be localised in the ER as well as in the plasma membrane (Castro J, Aromataris EC, Rychkov, G & Barritt GJ, unpublished results). In liver cells expressing TRPV1, the amount of Ca2+ released from the ER by a TRPV1 agonist (RTX), measured using the cytoplasmic fluorescent Ca2+ reporter fura-2, was found to be the same as that released by a SERCA inhibitor (DBHQ), indicating that TRPV1 agonist-sensitive stores substantially overlap with SERCA inhibitor-sensitive stores (results not shown). However, in contrast to SERCA inhibitors, TRPV1 agonists did not activate Ca2+ entry measured using fura-2 or patch clamp recording (Castro J, Aromataris EC, Rychkov, G & Barritt GJ, unpublished results). In cells expressing TRPV1, the release of Ca2+ from the ER could readily be detected using fura-2, but could not be detected using the fluorescent Ca2+ reporter FFP-18, which detects increases in intracellular Ca2+ concentration beneath the plasma membrane58 (Figure 1A,B). By contrast, Ca2+ release caused by SERCA inhibitors could be detected by both fura-2 and FFP-18 (results not shown). Taken together these results indicate that in cells expressing TRPV1, the region of the ER from which TRPV1 agonists release Ca2+ is some distance from the plasma membrane.

Figure 1

Figure 1. Ca2+ release from the ER of H4-IIE liver cells induced by the TRPV1 agonist resiniferatoxin (RTX) is detected by the cytoplasmic Ca2+ reporter fura-2 (A) but not by the near plasma membrane intracellular Ca2+ reporter FFP-18 (B), whereas Ca2+ release induced by the choleretic bile acid TDCA is not detected by fura-2 (C) but is detected by FFP-18 (D). A: In the absence of Ca2+ext, RTX induces the release of Ca2+ from the ER in H4-IIE cells ectopically expressing TRPV1 (TRPV1(+) cells) but not in cells which do not express TRPV1 (TRPV1(-)) cells. Subsequent addition of DBHQ releases no further Ca2+ in TRPV1(+) cells but does release Ca2+ in the TRPV1(-) cells. B: In cells loaded with FFP-18 and incubated in the absence of Ca2+ext (after prior incubation at 1.3 mM Ca2+ext), RTX does not cause a detectable increase in [Ca2+]cyt. Subsequent addition of Ca2+ext does lead to an increase in [Ca2+]cyt. C: In cells loaded with fura-2 incubated in the absence of Ca2+ext, taurocholic acid (TDCA) (300 μM) does not induce any detectable increase in [Ca2+]cyt, while the subsequent addition of 10 μM DBHQ causes a substantial transient increase in [Ca2+]cyt. D: In cells loaded with FFP-18 incubated in the absence of Ca2+ext, TDCA induces a detectable increase in [Ca2+]cyt. H4-IIE cells were loaded with fura-2 or FFP-18 and changes in [Ca2+]cyt were measured by confocal fluorescence microscopy. The times of addition of reagents are indicated by the horizontal bars (from Castro J, Aromataris EC, Rychkov, G & Barritt GJ, unpublished results).

Figure 2

Figure 2. In contrast to SERCA inhibitor thapsigargin (Tg), the TRPV1 agonist RTX does not induce a redistribution of STIM1 in H4-IIE liver cells ectopically expressing TRPV1, whereas taurodeoxycholic acid (TDCA) does induce a redistribution of STIM1 under experimental conditions similar to those where no TDCA-induced increase in [Ca2+]cyt is detected with fura-2 (cf. Figure 1C). A: Images obtained by confocal fluorescence microscopy of the distribution of STIM1, observed using STIM1-Cherry (STIM1), and the ER, observed using a YFP-tagged ER marker (ER-YFP) in the same cell, and the merged image (Merged). The STIM1-Cherry construct is a fluorescent reporter constructed by inserting the fluorescent mCherry protein after the signal sequence of hSTIM1.63 B: Cells ectopically expressing TRPV1 and STIM1-Cherry treated with vehicle (Control), 1 μM thapsigargin for 10 min (Tg), or 1 μM RTX for 10 min (RTX). C,D: TRPV1(+) cells expressing STIM1-Cherry were treated with 1 μM thapsigargin (Tg) (C) or 300 μM TDCA (TDCA) (D). The time elapsed (seconds) after addition of agonist is shown at the bottom of each frame. The scale bars represent 10 μm (from Castro J, Rychkov, G & Barritt GJ, unpublished results).

In liver cells incubated in the absence of agonist, STIM1 is distributed in the ER, as shown by the fluorescence images of STIM1-Cherry and YFP-tagged ER tracker (ER-YFP) in Figure 2A. In contrast to the effect of the SERCA inhibitor thapsigargin (Figure 2B middle panel (Tg) cf. Figure 2B left panel (control)) the TRPV1 agonist RTX did not cause a redistribution of STIM1 (Figure 2B right-hand panel (RTX) cf. Figure 2B left-hand panel (control)). In cells expressing TRPV1, incubated at zero Ca2+ext, the release of Ca2+ from the ER induced by TDCA could be detected by FFP-18 but not by fura-2 (Figure 1D cf. Figure 1C). Moreover, in TRPV1-expressing cells incubated at zero Ca2+ext, TDCA caused a redistribution of STIM1 to puncta similar to that caused by the SERCA inhibitor thapsigargin (Figure 2D (TDCA) cf. Figure 2C (Tg)). These results have provided further evidence that in liver cells Ca2+ release from a small component of the ER, which is presumably located near the plasma membrane, is required to induce STIM1 redistribution and SOC activation (Castro J, Aromataris EC, Rychkov, G & Barritt GJ, unpublished results).

Figure 3

Figure 3. A schematic representation of the proteins and organelles thought to be involved in the activation of store-operated Ca2+ channels in liver cells. The proposed mechanism of activation of SOCs can be summarized as follows. SOC activation requires a decrease in Ca2+ in the lumen of the ER in a subregion of the ER which is in close proximity to the plasma membrane and which forms ER-plasma membrane junctions. The ER subregion is enriched in type 1 IP3Rs. While the ER subregion communicates with the bulk of the ER, the movement of Ca2+ between the subregion and the bulk of the ER is slow. The steps in the activation of SOCs are: the initiating decrease in [Ca2+] in the lumen of the ER induced by IP3 (physiological) or a SERCA inhibitor (experimental), dissociation of Ca2+ from the luminal domain of the Ca2+ sensor STIM1, a conformational change in STIM1, oligomerisation of STIM1, relocalisation of STIM1 in the ER, interaction of STIM1 in close proximity to ER-plasma membrane junctions with CRACM1/Orai1, leading to a conformational change and increase the probability of opening of the Orai1 channel. Other proteins (as yet unidentified) are likely to be involved. The structure of the F-actin cytoskeleton, regulated in part by Gi2α and PLCγ1, is thought to play permissive roles in the activation pathway. Ca2+ which moves through SOCs into the ER-plasma membrane junction may cause a local increase in [Ca2+]cyt at the mouth of the channel, before being transported directly to the lumen of the ER via SERCA pumps, and to mitochondria (adopted from Barritt et al., 20083).

Ca2+ microdomains and the regulation of Ca2+ entry through liver cell SOCs

Current ideas for the mechanism of activation of SOCs in liver cells are summarized in Figure 3. This shows in schematic form the proposed subregion of the ER which is enriched in IP3R and located in ER-plasma membrane junctions, the roles of the Orai1 and STIM1 proteins as plasma membrane channel pore and ER Ca2+ sensor, respectively, and the proposed permissive roles of PLCγ1, Gi2 and F-actin. The activation mechanism involves several microdomains of intracellular Ca2+ in the cytoplasmic space and in the ER. The results obtained using ectopically-expressed TRPV1 and TDCA, described above, suggest that Ca2+ release from the bulk of the ER is not required, or at least is not critical, for SOC activation. It is proposed that the essential component of the ER, as far as SOC activation is concerned, is a putative ER subregion which is enriched in IP3R1 and is presumably located at ER-plasma membrane junctions. It is implied, but yet to be tested directly, that luminal movement of Ca2+ between the bulk of the ER and this ER subregion close to the plasma membrane is slow relative to the timescale of SOC activation. Further, that another property of the ER subregion which differentiates it from the bulk of the ER is that ectopically-expressed TRPV1 is not localised in this subregion.

Results obtained from studies with other cell types indicate that the activation of SOCs and regulation of the flow of Ca2+ through SOCs involves at least one microdomain of Ca2+ in the cytoplasmic space. This is a local increase in Ca2+ in the ER-plasma membrane junctional space at the mouth of the SOC (Orai) channel which occurs after channel activation. This, in part, is responsible for feedback inhibition of the channel itself, and may be responsible for the regulation of some enzymes such as adenylate cyclase.4,59-62 The transport of Ca2+ from this putative microdomain to the ER and mitochondria plays an important role, not only in refilling the ER Ca2+ stores, but also in regulating the feedback inhibition by Ca2+ of the SOC channel (reviewed by Parekh & Putney4). Interpretation of some results obtained in studies investigating the role of Ca2+ in this microdomain in regulating the activities of enzymes such as adenylate cyclase is complex as often the experiments were conducted in the presence of a SERCA inhibitor (e.g. thapsigargin) which would cause a much larger increase in Ca2+ in this microdomain than would occur under physiological conditions.59

Conclusions

It can be concluded that liver cells express SOCs with a high selectivity for Ca2+ and with properties essentially similar to those of CRAC channels in lymphocytes and mast cells. Orai polypeptides and STIM1 polypeptides constitute the pore and Ca2+ sensor of the liver cell SOC, respectively. The activation mechanism involves Ca2+ release from a putative small subregion of the ER which is enriched in IP3R1 and likely close to the plasma membrane. Further experiments might be directed to investigating the nature of the ER subregion and its relationship with the bulk of the ER, especially in connection with the steps involved in STIM1 movement, and oligomerisation, and the interaction of STIM1 with Orai1 and other proteins.

Acknowledgements

The authors gratefully acknowledge the assistance of Diana Kassos in the preparation of the manuscript. Research conducted in the authors’ laboratories which has contributed to this review is supported by grants from the National Health and Medical Research Council of Australia, the Australian Research Council, and the Flinders Medical Centre Foundation of South Australia.

References

1.  Berridge MJ, Bootman MD, Roderick HL. Calcium signalling: dynamics, homeostasis and remodelling. Nature Rev. Mol. Cell Biol. 2003; 4: 517-29.

2.  Woods NM, Cuthbertson KS, Cobbold PH. Repetitive transient rises in cytoplasmic free calcium in hormone-stimulated hepatocytes. Nature 1986; 319: 600-2.

3.  Barritt GJ, Chen J, Rychkov, G. Ca2+-permeable channels in the hepatocyte plasma membrane and their roles in hepatocyte physiology. Biochim. Biophys. Acta Mol. Cell Res. 2008; 1783: 651-72.

4.  Parekh AB, Putney JW Jr. Store-operated calcium channels. Physiol. Rev. 2005; 85: 757-810.

5.  Cao Y, Chatton JY. Involvement of calmodulin in the activation of store-operated Ca2+ entry in rat hepatocytes. FEBS Lett. 1998; 424: 33-6.

6.  Carini R, Bellomo G, Paradisi L, Dianzani MU, Albano E. 4-Hydroxynonenal triggers Ca2+ influx in isolated rat hepatocytes. Biochem. Biophys. Res. Commun. 1996; 218: 772-6.

7.  Guihard G, Noel J, Capiod T. Ca2+ depletion and inositol 1,4,5-trisphosphate-evoked activation of Ca2+ entry in single guinea pig hepatocytes. J. Biol. Chem. 2000; 275: 13411-4.

8.  Striggow F, Bohnensack R. Inositol 1,4,5-trisphosphate activates receptor-mediated calcium entry by two different pathways in hepatocytes. Eur. J. Biochem. 1994; 222: 229-34.

9.  Hansen CA, Yang LJ, Williamson JR. Mechanisms of receptor-mediated Ca2+ signaling in rat hepatocytes. J. Biol. Chem. 1991; 266: 18573-9.

10.  Ikari A, Sakai H, Takeguchi N. ATP, thapsigargin and cAMP increase Ca2+ in rat hepatocytes by activating three different Ca2+ influx pathways. Jap. J. Physiol. 1997; 47: 235-9.

11.  Junankar PR, Karjalainen A, Kirk K. The role of P2Y1 purinergic receptors and cytosolic Ca2+ in hypotonically activated osmolyte efflux from a rat hepatoma cell line. J. Biol. Chem. 2002; 277: 40324-34.

12.  Kass GE, Llopis J, Chow SC, Duddy SK, Orrenius S. Receptor-operated calcium influx in rat hepatocytes. Identification and characterization using manganese. J. Biol. Chem. 1990; 265: 17486-92.

13.  Llopis J, Kass GE, Gahm A, Orrenius S. Evidence for two pathways of receptor-mediated Ca2+ entry in hepatocytes. Biochem. J. 1992; 284: 243-7.

14.  Zhang Y, Duszynski J, Hreniuk S, Waybill MM, LaNoue KF. Regulation of plasma membrane permeability to calcium in primary cultures of rat hepatocytes. Cell Calcium 1991; 12: 559-75.

15.  Fernando KC, Barritt GJ. Characterisation of the inhibition of the hepatocyte receptor-activated Ca2+ inflow system by gadolinium and SK&F 96365. Biochim. Biophys. Acta 1994; 1222: 383-9.

16.  Applegate TL, Karjalainen A, Bygrave FL. Rapid Ca2+ influx induced by the action of dibutylhydroquinone and glucagon in the perfused rat liver. Biochem. J. 1997; 323: 463-7.

17.  Altin JG, Bygrave FL. The influx of Ca2+ induced by the administration of glucagon and Ca2+-mobilizing agents to the perfused rat liver could involve at least two separate pathways. Biochem. J. 1987; 242: 43-50.

18.  Rychkov G, Brereton HM, Harland ML, Barritt GJ. Plasma membrane Ca2+ release-activated Ca2+ channels with a high selectivity for Ca2+ identified by patch-clamp recording in rat liver cells. Hepatology 2001; 33: 938-47.

19.  Litjens T, Harland ML, Roberts ML, Barritt GJ, Rychkov GY. Fast Ca2+-dependent inactivation of the store-operated Ca2+ current (ISOC) in liver cells: a role for calmodulin. J. Physiol. 2004; 558: 85-97.

20.  Rychkov GY, Litjens T, Roberts ML, Barritt GJ. ATP and vasopressin activate a single type of store-operated Ca2+ channel, identified by patch-clamp recording, in rat hepatocytes. Cell Calcium 2005; 37: 183-91.

21.  Gregory RB, Rychkov G, Barritt G J. Evidence that 2-aminoethyl diphenylborate is a novel inhibitor of store-operated Ca2+ channels in liver cells, and acts through a mechanism which does not involve inositol trisphosphate receptors. Biochem. J. 2001; 354: 285-90.

22.  Auld A, Chen J, Brereton HM, Wang Y-J, Gregory RB, Barritt GJ. Store-operated Ca2+ inflow in Reuber hepatoma cells is inhibited by voltage-operated Ca2+ channel antagonists and, in contrast to freshly isolated hepatocytes, does not require a pertussis toxin-sensitive trimeric GTP-binding protein. Biochim. Biophys. Acta 2000; 1497: 11-26.

23.  Rychkov GY, Litjens T, Roberts ML, Barritt GJ. Arachidonic acid inhibits the store-operated Ca2+ current in rat liver cells. Biochem. J. 2005; 385: 551-6.

24.  Berven LA, Barritt GJ. Evidence obtained using single hepatocytes for inhibition by the phospholipase C inhibitor U73122 of store-operated Ca2+ inflow. Biochem. Pharmacol. 1995; 49: 1373-9.

25.  Litjens T, Nguyen T, Castro J, Aromataris EC, Jones L, Barritt GJ, Rychkov GY. Phospholipase C-γ1 is required for the activation of store-operated Ca2+ channels in liver cells. Biochem. J. 2007; 405: 269-76.

26.  Litjens T, Harland ML, Roberts ML, Barritt GJ, Rychkov GY. Fast Ca2+-dependent inactivation of the store-operated Ca2+ current (ISOC) in liver cells: a role for calmodulin. J. Physiol. 2004; 558: 85-97.

27.  Lewis RS. The molecular choreography of a store-operated calcium channel. Nature 2007; 446: 284-7.

28.  Liou J, Fivaz M, Inoue T, Meyer T. Live-cell imaging reveals sequential oligomerization and local plasma membrane targeting of stromal interaction molecule 1 after Ca2+ store depletion. Proc. Nat. Acad. Sci. USA 2007; 104: 9301-6.

29.  Varnai P, Toth B, Toth DJ, Hunyady L, Balla T. Visualization and manipulation of plasma membrane-endoplasmic reticulum contact sites indicates the presence of additional molecular components within the STIM1-Orai1 complex. J. Biol. Chem. 2007; 282: 29678-90.

30.  Hogan PG, Rao A. Dissecting ICRAC, a store-operated calcium current. Tr. Biochem. Sci. 2007; 32: 235-45.

31.  Smyth JT, Dehaven WI, Bird GS, Putney JW Jr. Ca2+-store-dependent and -independent reversal of Stim1 localization and function. J. Cell Sci. 2008; 121: 762-72.

32.  Aromataris EC, Castro J, Rychkov G, Barritt GJ. Store-operated Ca2+ channels and Stromal Interaction Molecule 1 (STIM1) are targets for the actions of bile acids on liver cells. Biochim. Biophys. Acta Mol. Cell Res. 2008; 1783: 874-85.

33.  Lievremont J-P, Hill A-M, Hilly M, Mauger J-P. The inositol 1,4,5-trisphosphate receptor is localised on specialised sub-regions of the endoplasmic reticulum in rat liver. Biochem. J. 1994; 300: 419-27.

34.  Lievremont JP. et al. Intracellular calcium stores and inositol 1,4,5-trisphosphate receptor in rat liver cells. Biochem. J. 1996; 314: 189-97.

35.  Rychkov G, Barritt GJ. TRPC1 Ca2+-permeable channels in animal cells. Handb. Exp. Pharmacol. 2007; 179: 23-52.

36.  Yuan JP, Zeng W, Huang GN, Worley PF, Muallem S. STIM1 heteromultimerizes TRPC channels to determine their function as store-operated channels. Nature Cell Biol. 2007; 9: 636-45.

37.  Brereton HM, Chen J, Rychkov G, Harland ML, Barritt GJ. Maitotoxin activates an endogenous non-selective cation channel and is an effective initiator of the activation of the heterologously expressed hTRPC-1 (transient receptor potential) non-selective cation channel in H4-IIE liver cells. Biochim. Biophys. Acta 2001; 1540: 107-26.

38.  Chen J, Barritt GJ. Evidence that TRPC1 (transient receptor potential canonical 1) forms a Ca2+-permeable channel linked to the regulation of cell volume in liver cells obtained using small interfering RNA targeted against TRPC1. Biochem. J. 2003; 373: 327-36.

39.  Ramsey IS, Delling M, Clapham DE. An introduction to TRP channels. Ann. Rev. Physiol. 2006; 68: 619-47.

40.  Nilius B, Owsianik G, Voets T, Peters JA. Transient receptor potential cation channels in disease. Physiol. Rev. 2007; 87: 165-217.

41.  Berven LA, Hughes BP, Barritt GJ. A slowly ADP-ribosylated pertussis-toxin-sensitive GTP-binding regulatory protein is required for vasopressin-stimulated Ca2+ inflow in hepatocytes. Biochem. J. 1994; 299: 399-407.

42.  Butta N, Urcelay E, Gonzalez-Manchon C, Parrilla R, Ayuso MS. Pertussis toxin inhibition of α1-adrenergic or vasopressin-induced Ca2+ fluxes in rat liver. Selective inhibition of the α1-adrenergic receptor-coupled metabolic activation. J. Biol. Chem. 1993; 268: 6081-9.

43.  Berven LA, Barritt GJ. A role for a pertussis toxin-sensitive trimeric G-protein in store-operated Ca2+ inflow in hepatocytes. FEBS Lett. 1994; 346: 235-40.

44.  Fernando KC, Barritt GJ. Evidence from studies with hepatocyte suspensions that store-operated Ca2+ inflow requires a pertussis toxin-sensitive trimeric G-protein. Biochem. J. 1994; 303: 351-6.

45.  Berven LA, Crouch MF, Katsis F, Kemp BE, Harland LM, Barritt GJ. Evidence that the pertussis toxin-sensitive trimeric GTP-binding protein Gi2 is required for agonist- and store-activated Ca2+ inflow in hepatocytes. J. Biol. Chem. 1995; 270: 25893-7.

46.  Wang YJ, Gregory RB, Barritt GJ. Regulation of F-actin and endoplasmic reticulum organization by the trimeric G-protein Gi2 in rat hepatocytes. Implication for the activation of store-operated Ca2+ inflow. J. Biol. Chem. 2000; 275: 22229-37.

47.  Wang YJ, Gregory RB, Barritt GJ. Maintenance of the filamentous actin cytoskeleton is necessary for the activation of store-operated Ca2+ channels, but not other types of plasma-membrane Ca2+ channels, in rat hepatocytes. Biochem. J. 2002; 363: 117-26.

48.  Fernando KC, Gregory RB, Katsis F, Kemp BE, Barritt GJ. Evidence that a low-molecular-mass GTP-binding protein is required for store-activated Ca2+ inflow in hepatocytes. Biochem. J. 1997; 328: 463-71.

49.  Koelle MR. Heterotrimeric G protein signaling: Getting inside the cell. Cell 2006; 126: 25-7.

50.  Schaefer M, Petronczki M, Dorner D, Forte M, Knoblich JA. Heterotrimeric G proteins direct two modes of asymmetric cell division in the Drosophila nervous system. Cell 2001; 107: 183-94.

51.  Hirata K, Pusl T, O'Neill AF, Dranoff JA, Nathanson MH. The type II inositol 1,4,5-trisphosphate receptor can trigger Ca2+ waves in rat hepatocytes. Gastroenterology 2002; 122: 1088-100.

52.  Lilly LB, Gollan JL. Ryanodine-induced calcium release from hepatic microsomes and permeabilized hepatocytes. Am. J. Physiol. 1995; 268, G1017-24.

53.  Wojcikiewicz RJ. Type I, II, and III inositol 1,4,5-trisphosphate receptors are unequally susceptible to down-regulation and are expressed in markedly different proportions in different cell types. J. Biol. Chem. 1995; 270: 11678-83.

54.  Hernandez E. et al. The spatial distribution of inositol 1,4,5-trisphosphate receptor isoforms shapes Ca2+ waves. J. Biol. Chem. 2007; 282: 10057-67.

55.  Gregory RB, Wilcox RA, Berven LA, van Straten NCR, van der Marel GA, van Boom JH, Barritt GJ. Evidence for the involvement of a small subregion of the endoplasmic reticulum in the inositol trisphosphate receptor-induced activation of Ca2+ inflow in rat hepatocytes. Biochem. J. 1999; 341: 401-8.

56.  Gregory RB, Hughes R, Riley AM, Potter BVL, Wilcox RA, Barritt, G.J. Inositol trisphosphate analogues selective for types I and II inositol trisphosphate receptors exert differential effects on vasopressin-stimulated Ca2+ inflow and Ca2+ release from intracellular stores in rat hepatocytes. Biochem. J. 2004; 381: 519-26.

57.  Rossier MF, Bird GS, Putney JW Jr. Subcellular distribution of the calcium-storing inositol 1,4,5-trisphosphate-sensitive organelle in rat liver. Possible linkage to the plasma membrane through the actin microfilaments. Biochem. J. 1991; 274: 643-50.

58.  Davies EV, Blanchfield H, Hallett MB. Use of fluorescent dyes for measurement and localization of organelles associated with Ca2+ store release in human neutrophils. Cell Biol. Internat. 1997; 21: 655-63.

59.  Chan C. et al. Evaluation, using targeted aequorins, of the roles of the endoplasmic reticulum and its (Ca2++Mg2+)ATP-ases in the activation of store-operated Ca2+ channels in liver cells. Cell Calcium 2004; 35: 317-31.

60.  Chiono M, Mahey R, Tate G, Cooper D.M. Capacitative Ca2+ entry exclusively inhibits cAMP synthesis in C6-2B glioma cells. Evidence that physiologically evoked Ca2+ entry regulates Ca2+-inhibitable adenylyl cyclase in non-excitable cells. J. Biol. Chem. 1995; 270: 1149-55.

61.  Fagan KA, Smith KE, Cooper DM. Regulation of the Ca2+-inhibitable adenylyl cyclase type VI by capacitative Ca2+ entry requires localization in cholesterol-rich domains. J. Biol. Chem. 2000; 275: 26530-7.

62.  Liou J, Fivaz M, Inoue T, Meyer T. Live-cell imaging reveals sequential oligomerization and local plasma membrane targeting of stromal interaction molecule 1 after Ca2+ store depletion. Proc. Nat. Acad. Sci. USA 2007; 104: 9301-6.

63.  Luik RM, Wu MM, Buchanan J, Lewis RS. The elementary unit of store-operated Ca2+ entry: local activation of CRAC channels by STIM1 at ER-plasma membrane junctions. J. Cell Biol. 2006; 174: 815-25.


Received 2 July 2008, in revised form 11 September 2008. Accepted 11 September 2008.
© G.J. Barritt 2008.