Abstract
Pancreatitis and Calcium Signalling was an international research workshop organized by the authors and held at the Liverpool Medical Institution, Liverpool, United Kingdom, from Sunday 12th to Tuesday 14th November 2006. The overall goal of the workshop was to review progress and explore new opportunities for understanding the mechanisms of acute pancreatitis with an emphasis on the role of pathological calcium signaling. The participants included those with significant interest and expertise in pancreatitis research and others who are in fields outside gastroenterology but with significant expertise in areas of cell biology relevant to pancreatitis. The workshop was designed to enhance interchange of ideas and collaborations, to engage and encourage younger researchers in the field, and promote biomedical research through the participating and supporting organizations and societies. The workshop was divided into 8 topic-oriented sessions. The sessions were: (1) Physiology and pathophysiology of calcium signaling; (2) Interacting signaling mechanisms; (3) Premature digestive enzyme activation; (4) Physiology Society Lecture: Aberrant Ca2+ signaling, bicarbonate secretion, and pancreatitis; (5) NFJB, cytokines, and immune mechanisms; (6) Mitochondrial injury; (7) Cell death pathways; and (8) Overview of areas for future research. In each session, speakers presented work appropriate to the topic followed by discussion of the material presented by the group.
The publication of these proceedings is intended to provide a platform for enhancing research and therapeutic development for acute pancreatitis.
PHYSIOLOGICAL AND PATHOLOGICAL CA2+ SIGNALING
Overview of Acinar Cell Ca2+ Signaling
Ole Petersen (Liverpool, UK) began by acknowledging the major contribution to understanding cell signaling made by patch-clamp technology, introduced by the 1991 Nobel laureates Bert Sakmann and Erwin Neher for their discoveries concerning the function of single ion channels in cells.1 After secretagogue receptor activation, the second messengers inositol trisphosphate (IP3), cyclic adenosine diphosphate ribose (cADPR), and nicotinic acid adenine dinucleotide phosphate (NAADP) all induce release of Ca2+ through their respective receptor (IP3R and RyR) Ca2+ channels from internal stores into the apical cytosol (Fig. 1).1–5 The Ca2+ waves can only be initiated within the apical pole and are usually confined there by the perigranular mitochondrial buffer zone, where the Ca2+ signals initiate stimulus-metabolism coupling. The Ca2+ signals originate from terminals of the endoplasmic reticulum (ER) (via IP3, NAADP, and cADPR) and an acidic store that recent data indicate includes both the lysosomes (via NAADP) and zymogen granules (ZGs) (via IP3 and cADPR).2,3 Globalization of Ca2+ signals results from combinatorial action of the 3 second messengers, as well as Ca2+-induced Ca2+ release, which acts particularly on the RyR. All Ca2+ signals are the result of the interaction of second messenger–mediated release, which produces agonist-specific signatures. Second-messenger effects from acetylcholine (ACh) stimulation are largely mediated through IP3-mediated Ca2+ release, whereas those from cholecystokinin (CCK) are predominantly through NAADP and cADPR. When adenosine triphosphate (ATP) supplies fail, Ca2+ cannot be cleared from the cytosol into the ER by the sarcoendoplasmic reticulum ATPase Ca2+ pump (SERCA) or out of the cell through the plasma membrane ATPase Ca2+ pump (PMCA).4 This is especially dangerous for the pancreatic acinar cell, which lacks a Na+/Ca2+ exchanger.1,5
FIGURE 1.
Ca2+ release mechanisms in the pancreatic acinar cell. A, Cross-sectional representation of Ca2+ pools, with the secretory pole to left (high sensitivity) where cytosolic Ca2+ signals begin, in response to the second messengers IP3, cyclic ADP ribose, or NAADP acting on their calcium channel receptors (IP3R, RyR, and/or NAADPR). Experimental application (via patch pipette) of a single second messenger (IP3 or RyR or NAADPR) elicits Ca2+ signals localized to the apical pole by the perigranular mitochondrial firewall. B, Combinations of second messengers evoke global cytosolic Ca2+ signals that spread beyond the mitochondrial firewall throughout the cell (into the low-sensitivity basolateral area), amplified by Ca2+-induced Ca2+ release. Ca2+-ATPase pumps clear Ca2+ from the cytosol back into the ER and out of the cell. Toxic Ca2+ release occurs when the cytosol is overloaded and Ca2+ is not cleared, for example, as a result of mitochondrial failure. C, Ca2+ release mechanisms in the high sensitivity apical pole, where ER, endolysosomal and ZG Ca2+ stores interact. IP3 and cADPR elicit Ca2+ release from ER terminals and ZGs, whereas NAADP elicits Ca2+ release particularly from lysosomes, with amplification by Ca2+-induced Ca2+ release. Different second messengers predominate after stimulation by different secretagogues, for example, IP3 after ACh, NAADP after CCK.
In the discussion of Dr Petersen’s presentation, it was emphasized that Ca2+ influx is as important for maintenance of signal generation as is release of Ca2+ from the ER at least in part because the Ca2+ influx is necessary for recharging the ER with calcium.
Ca2+ Entry in Health and Disease
Anant Parekh (Oxford, United Kingdom) introduced store-operated Ca2+ entry channels (SOCs), a family of plasma membrane Ca2+ channels that open in response to active (eg, second messenger–mediated) or passive (eg, SERCA pump inhibition) ER Ca2+ depletion.6 The best described and most widely distributed is the Ca2+ release–activated Ca2+ channel (CRAC), characterized electrophysiologically by patch-clamp measurements of the whole-cell Ca2+ current, ICRAC. The CRAC channels are critical not only to store refilling, but also to ATP production, exocytosis, and cell proliferation. Furthermore, CRAC channel abnormalities are implicated in human disease. The ICRAC is highly selective for Ca2+, with 10,000 times greater relative permeability to that of Na+; single-channel conductance is low, and of the order of 8000 CRAC, channels are estimated to be present in the plasma membrane of a single cell. RNA interference knockdown approach has demonstrated that stromal interaction molecule 1 (STIM-1), which spans the ER membrane, is a sensor that upon ER Ca2+ depletion migrates to puncta below the plasma membrane to trigger CRAC opening. The Orai1/CRACM1 is a plasma membrane protein, the selectivity of which to Ca2+ is changed by site-directed mutagenesis, revealing that Orai1 is the CRAC channel pore.7 To simulate physiological CRAC channel opening in isolated cells, mitochondria are energized with a cocktail of metabolic substrates, by means of which they become hyperpolarized. This has permitted demonstration that mitochondria activate the CRAC channel and reduce Ca2+-dependent inactivation of ICRAC.8 In addition, paired patch-clamp recordings from 2 cells in close proximity have demonstrated that a cytoplasmic rise in Ca2+ after ICRAC activation in one cell results in a paracrine signal from that cell which activates ICRAC in the second cell. Subsequent work on rat mast cells has shown that this signal is leukotriene C4 (LTC4), the release of which follows the initiation of a signaling cascade by Ca2+ via protein kinase C (PKC) α and β1, cytosolic phospholipase A2, the mitogen-activated protein kinase MEK/ERK pathway and then 5-lipoxygenase, converting arachidonic acid into LTC4.9,10 The relevance of this pathway to human disease is currently under exploration in human mast cells extracted from tissues affected by allergy.
The discussion of Dr Parekh’s presentation added that coordination of Ca2+ entry occurs via receptor-mediated mechanisms rather than gap junctions.
Critical Subcellular Compartments
Rosario Rizzuto (Ferrara, Italy) described how the transfection of targeted aequorin and of ATP luciferase had identified rapid large increases in mitochondrial Ca2+, which occur immediately after cytoplasmic Ca2+ signals, and which regulate mitochondrial function.11,12 When accompanied by oxidative and/or toxic stress, these changes induce cell death,13 discussed later (see Mitochondrial Injury later). The regulatory mechanisms are not all clear because the molecular identities of the Ca2+ uniporter, Na+/Ca2+ exchanger and H+/Ca2+ exchanger on the inner mitochondrial membrane remain unknown. Close contacts between the ER and mitochondria have been identified and characterized.14,15 Again using targeted aequorins and Ca2+-sensitive green fluorescent proteins, the voltage-dependent anion channel (VDAC) was found to be an essential part of this closely interacting micro-domain, allowing rapid transfer of Ca2+ from the ER through the outer mitochondrial membrane, which then enters the matrix through the Ca2+ uniporter.15 VDAC overexpression in HeLa cells was found to enhance mitochondrial Ca2+ uptake, confirming that transfer of Ca2+ from the mouth of the IP3R on the ER membrane into mitochondria is critically dependent on the VDAC repertoire. To find regulatory components of the VDAC complex, yeast 2-hybrid screens were conducted using VDAC as bait, identifying a strong interaction with the chaperone glucose-regulated protein 75 (GRP75, also known as mitochondrial heat shock protein 70 [HSP70] or mortalin). Recent work examining the effects on VDAC function of the outer mitochondrial membrane binding domain of the IP3R and GRP75 has shown that the binding domain of the IP3R increases Ca2+ entry through VDAC, and that coupling of the IP3R on the ER and VDAC on the outer mitochondrial membrane depends on the presence of GRP75, further defining the VDAC complex.
During discussion of the presentation, Dr Rizzuto stated that the capacity of mitochondria to cope with Ca2+ is considerable, but matters go awry when combined with another stress, suggesting that Ca2+ is permissive to injury. Further discussion raised the possibility that mitochondria have a role by removing Ca2+ from the cytosol to prevent pathology, which could be one reason for their close interaction with the ER. Other discussions raised possibilities for roles for PKCs and HSPs in the regulation of VDAC. There was an open and unresolved debate as to whether different types of mitochondria exist in the same cell, or whether apparent differences arise from different signals and microdomains occurring within different parts of the same cell.
INTERACTING SIGNALING MECHANISMS
PKC, Alcohol, and Pancreatitis
Stephen Pandol discussed his group’s original findings that acinar cells produce and respond to cytokines, underscoring the role of inflammation in acinar cell injury and pancreatitis. The role of PKC isoforms remains a particularly important focus for their work.16–20 Both CCK hyperstimulation and tumor necrosis factor-α (TNF-α) activate nuclear factor-κB (NF-κB) via phosphatidylinositol-specific phospholipase C (PI-PLC, by CCK only) and phosphatidylcholine-specific PLC (PC-PLC, by both CCK and TNF-α), forming 1,2 diacylglycerol (nota bene activated PI-PLC also produces IP3), which activates PKCs that induce NF-κB activation.16,17 Dissection of the PKC isoforms responsible showed that it is the combination of the novel isoforms PKC-δ and PKC-ε that induces NF-κB activation. Further work showed that ethanol sensitizes the pancreas to inflammation from CCK stimulation by up-regulating PKC-ε because without ethanol, 100 pM CCK induced PKC-δ but not NF-κB, whereas with ethanol, 100 pM CCK induced both PKC-δ and PKC-ε, and subsequently activated NF-κB.18,19 Very recent work has implicated PKCs in pathological basolateral exocytosis mediated by the syntaxin-binding protein Munc18c, consequent upon ethanol and CCK stimulation of pancreatic acinar cells.20
The discussion of the presentation by Dr Pandol indicated that it is not known whether PKC activation alone or Ca2+ release alone was sufficient to induce acinar cell injury. Furthermore, the effects on PKC activation of secretin, vasoactive intestinal polypeptide, JMV180 (with or without ethanol), or bile salts have not been determined. Nor was it known whether the effects of PKC were structural or enzymatic, but fatty acid ethyl esters (FAEEs) and acetaldehyde had been found to down-regulate PKC.
Phosphatidylinositide 3 Kinases in Pancreatitis
Anna Gukovskaya (Los Angeles, Calif) outlined relationships among phosphoinositides and the 3 classes of mammalian phosphatidylinositide 3 kinases (PI3Ks), classes IA and IB, II, and III, which phosphorylate by inducible (class I and II) and constitutive (class III) means. Receptor tyrosine kinases and G protein–coupled receptors initiate class I PI3K activation of the AKT (protein kinase B [PKB], a serine-threonine protein kinase belonging to the protein kinase A, G, and C superfamily) signaling pathway that has complex effects on protein synthesis, metabolism, and cell survival and cycling. Class III PI3K also contributes to vesicle sorting and autophagy. Whereas class I PI3K activation of AKT occurs via phosphorylation of phosphatidylinositol-4,5-bisphosphate (hydrolysis of which by PI-PLC forms IP3 and diacylglycerol) to phosphatidylinositol-3,4,5-trisphosphate, phosphatase and tensin homologue deleted on chromosome ten (PTEN) (a tumor-suppressor gene) inhibits AKT by reversing this reaction. The PTEN itself is inhibited by oxidation, which destabilizes PTEN, and phosphorylation, which stabilizes PTEN. The role of many signaling molecules in pancreatitis has been investigated using CCK hyperstimulation, sometimes supplemented with ethanol administration.21,22 Previous work with this model had not implicated class I PI3Ks in the pathogenesis of pancreatitis, but had shown that class III PI3K induces formation of phosphatidylinositol phosphate, which may impair lysosomal trafficking and contribute to trypsinogen activation. Further evidence for this effect was derived from the actions of wortmannin, a PI3K inhibitor, which reduced phosphatidylinositol phosphate levels and ameliorated both cerulein-induced and taurocholate-induced acute pancreatitis. Subsequent experiments, however, have indicated that AKT activation occurs early in the course of cerulein-induced acute pancreatitis, as well as during ethanol feeding, an effect attributed to PTEN inhibition. A more specific approach was adopted through the use of mice deficient in subunit 110γ of class IB PI3Kγ, implicated in signaling from G protein–coupled receptors in tissues that include the exocrine pancreas; both in cerulein-induced and choline-deficient ethionine-supplemented diet–induced acute pancreatitis, such mice developed less trypsinogen activation, pancreatic neutrophil infiltration, and acinar cell necrosis, with more acinar cell apoptosis.23 In addition, in contrast to the effects of wortmannin, lack of 110γ resulted in less NK-κB activation.
Significant parallel work has shown that PI3K regulates Ca2+ signaling in isolated pancreatic acinar cells through an inhibitory effect on SERCA after hyperstimulation24 or bile salt exposure25; knockout of PI3Kγ prevented cytosolic overloading of Ca2+ and trypsinogen activation with these toxins.24,25
Discussion of Dr Gukovskaya’s presentation revealed that in acute pancreatitis, the relative importance of AKT activation versus trypsinogen activation is not known. Furthermore, the mechanisms that relate PI3K and PTEN activity to trypsinogen activation are not known. It is of note, however, that both PI3K inhibition with wortmannin or LY-294002 and 110γ deletion reduced Ca2+ mobilization and influx, indicating that PI3K inhibits SERCA pump activity and promotes Ca2+ mobilization into the cytosol, which in conditions of stress (eg, hyperstimulation) would tend to induce acinar cell injury and trypsinogen activation. Elucidation of the roles of different PI3Ks should be a topic of future research.
Novel Mechanisms of Nitric Oxide Release in Pancreatic Acinar Cells
Alexei Tepikin (Liverpool, United Kingdom) presented work investigating the role of nitric oxide (NO) generation and nitrosative stress in pancreatic acinar cells,26,27 building on novel techniques using fluorescent measurement of subcellular responses to toxins and cell stress.28–30 Nitric oxide was found to be released in isolated mouse pancreatic acinar cells hyperstimulated with either 10 μM ACh or 10 nM CCK. Nitric oxide responses were also seen after more representative, quasiphysiological secretagogue concentrations (50 nM ACh, 10 pM CCK), although in fewer cells, so most experiments were conducted with hyperstimulation.27 This work was conducted using intracellular dialysis with whole-cell patch clamp, measuring intracellular NO release with the fluorescent dye 4-amino-5-methylamino-2′,7′-difluorofluorescein, most probably because this approach removed mobile intracellular nitrosation scavengers. 1,2-Bis(o-aminophenoxy) ethane-N,N,N′,N′-tetraacetic acid (BAPTA) inhibited ACh-induced responses of 4-amino-5-methylamino-2′,7′-difluorofluorescein, indicating that the NO release was dependent on cytosolic Ca2+ rises, and reduced glutathione and NO scavengers suppressed ACh-induced NO responses. Further analysis of the source of NO release, notably through the combination of HgCl2 (which catalyzes S-NO bond cleavage and has been used to quantify S-nitrosothiols) and ionomycin, showed intracellular S-nitrosothiols to be the primary source, with NO synthase (NOS) activity contributing to subsequent NO production. Interestingly, NO responses were found to be independent of calmodulin, PKC, and transition metals, but were inhibited by calpain antagonists, μ-calpain and m-calpain, having previously been established as present in the exocrine pancreas. Putatively, calpain-induced partial proteolysis of S-nitrosylated proteins may promote thiol denitrosylation.
Comment during the discussion of Dr Tepikin’s presentation indicated that, from evidence to date, NO does not appear important in the early acinar cell events of acute pancreatitis but may be important in vascular events. The pancreatic acinar cell contains all 3 types of NOS (neuronal NOS, endothelial NOS [eNOS], and inducible NOS), which are activated by Ca2+. Previous experiments show that during the first 8 hours of cerulein-induced acute pancreatitis, there is evidence of an increase in eNOS activity; and eNOS deletion impairs the increase in pancreatic blood flow that occurs during cerulein hyperstimulation, resulting in more severe pancreatitis.
PREMATURE DIGESTIVE ENZYME ACTIVATION
The Granule, an Intracellular Ion Oscillator: Implications for the Pathology of Secretion
Pedro Verdugo (Washington) explained that within secretory granules in many cell types, there is a polymer matrix network that behaves as an ion exchange resin and which determines how the signaling of secretion is coordinated. The polymer matrix (eg, heparin in mast cell granules, chromogranin in chromaffin cell granules) holds the secretory product (eg, histamine from mast cells, adrenaline from chromaffin cells), together with Ca2+ and H+ ions as well as ATP in the granules. The polymer matrix undergoes a phase transition, depending on local physicochemical factors, notably a change in granular Ca2+ concentration; with a small decrease in concentration, a large increase in volume occurs.31 The transition is dependent on the second power of the valence of this ion. Thus, during passage through the signal transduction domain of the apical secretory pole, granules hold secretory products tightly in their polymer matrix, but during exocytosis, Na+ and K+ replace Ca2+, and the matrix expands to release the secretory product.32 Image deconvolution studies of isolated granules from many cell types have shown oscillations induced by IP3 and cADPR through Ca2+, K+, and H+ channels in the membrane of granules, thought to prime effector mechanisms in secretion and related processes.33–35
The discussion of Dr Verdugo’s work centered on the amount of Ca2+ present in ZGs, and whether granular Ca2+ reloading occurs. The concentration of Ca2+ is 10 mM in the granular matrix and 50 μM in the granular solution; high levels of stimulation cause a drop in total granular Ca2+. In addition, the release of Ca2+ from the ZGs is followed by reloading; and Ca2+ oscillations occur and are probably dependent on reloading.
Determinants of Activation and Activity: Ca2+ Regulation of Trypsinogen Activation and Degradation
Miklos Sahin-Toth (Boston, Mass) focused on the effects of Ca2+ within the context of various interacting digestive enzymes that may contribute to, or protect from, acinar cell damage and thus sporadic or hereditary pancreatitis.36–40 The concentration of Ca2+ alters the speed at which trypsin activates trypsinogen (the rapid phase of the S-shaped curve of autoactivation) and degrades itself (as well as trypsinogen).38 Ca2+ does not, however, significantly affect the activation of trypsinogen by enterokinase, the activation of other zymogens by trypsin, or the digestion of dietary proteins by trypsin. Concentrations of Ca2+ in pancreatic juice and in the duodenum are normally greater than 1 mM, at which there is plentiful trypsin activity and robust trypsinogen activation, but little trypsin degradation. At lower Ca2+ concentrations (10–100 μM or less), autoactivation of human cationic trypsinogen (PRSS1, the major human trypsinogen) is less; notably, the Ca2+ binding site of trypsinogen (Glu75 to Glu 85) has a dissociation constant of 25 to 50 μM. Chymotrypsin C (formerly known as enzyme Y) accelerates trypsinogen activation through a Ca2+-independent effect on Asp218, which otherwise maintains an electrostatic interaction with the activation peptide.39 This effect of chymotrypsin C is further enhanced by the A16V mutation of the PRSS1 gene that causes about 5% of cases of hereditary pancreatitis. Chymotrypsin C also accelerates trypsin (and trypsinogen) degradation by cleaving the Ca2+ binding loop (Glu75 to Glu85), an effect that is inversely proportional to the Ca2+ concentration, near maximal at 25 μM and minimal at greater than 1 mM. The degradative action of chymotrypsin C is trypsin dependent because the R122H mutation of PRSS1 (at the trypsin cleavage site) that causes about 70% of cases of hereditary pancreatitis40 is resistant to its action, as is R122A produced by site-directed mutagenesis. If significant trypsinogen activation occurs within ZGs as a result of (in part) a loss of granular Ca2+, chymotrypsinogen C activation would be likely, leading to chymotrypsin C–enhanced degradation of trypsin, a potential line of defense against premature intracellular digestive enzyme activation.
During discussion of his presentation with the audience, Dr Sahin-Toth stated that there are no known mutations of chymotrypsin C, although this enzyme has been found in all mammalian pancreata examined, and it is more similar to elastase than trypsin.
Onset Events in the Apical Pole
Ashok Saluja (Minneapolis, Minn) advocated the view derived from many experimental data that pancreatitis begins with an acinar cell insult that blocks secretion, inducing colocalization of lysosomes and ZGs.41–45 These data include comparison of the effects of cerulein, which induces premature intracellular digestive enzyme activation, secretory block, apical-to-basolateral F-actin redistribution, and pancreatitis, with those of CCK-JMV180 and bombesin, which induce digestive enzyme activation without secretory block, F-actin redistribution, or pancreatitis.45 Admixture of the content of lysosomes and ZGs leads to trypsinogen activation by lysosomal cathepsin B, which if inhibited, markedly reduces trypsinogen activation, plasma amylase (prognostic in rodents), pancreatic edema, pancreatic necrosis, and myeloperoxidase activity. Results indicate that colocalization is a result of an abnormal elevation of the cytosolic Ca2+ concentration ([Ca2+]I) produced by a toxic insult, the effects of which can be prevented or reduced by Ca2+ chelators (eg, BAPTA), heat shock proteins, PI3K inhibitors (eg, low concentrations of wortmannin) or cathepsin B inhibition.42–44 The secretory block induced by hyperstimulation or L-arginine can be overcome by protease-activated receptor 2 activation, for example, with the peptide SLIGRL, which ameliorates pancreatitis, confirming the importance of secretory block in pathogenesis.
Further information shared during the discussion of Dr Saluja’s presentation brought up the point that the actin coating of ZGs may be upset by a persistently high [Ca2+]I, which may induce secretory block that could augment the zymogen activation in the cell. There are substantial data to support the view that trypsinogen activation occurs in condensing vacuoles, subsequently spreading elsewhere. Considering that there are two stages to secretion-fusion of granules with the apical plasma membrane followed by pore expansion, there are very elegant data showing that without pore expansion, no secretion occurs. The absence of pore expansion may be important in the pathogenesis of acute pancreatitis.
Other discussions focused on whether enzyme activation is the first event in acinar cell injury and the fact that colocalization has never been shown to occur before trypsinogen activation, a weakness in the colocalization hypothesis. It is not known how the pH at which cathepsin B activates trypsinogen is achieved in vivo, that is, with a maximum effect at pH 4.0 and no effect above pH 5.0. Nevertheless, 80% of pancreatic trypsinogen activation is lost during hyperstimulation-induced pancreatitis in the cathepsin B–knockout mouse, suggesting a direct effect, and overcoming a problem of experiments with CA-074me, which also inhibits cathepsin L, a lysosomal enzyme that degrades trypsinogen. The mechanism by which protease-activated receptor 2 activation restores secretion is not known.
Early Extracellular Events
Markus Lerch (Griefswald, Germany) provided a perspective with rhetorical questions and answers, complementing the focus of the workshop on intracellular events, notably Ca2+46 and cathepsin B,47 up to this point. The extravasation of fluid and recruitment of leukocytes, notably polymorphonuclear (PMN) leukocytes (neutrophils), are important early events in acute pancreatitis associated with a breakdown of intercellular adhesion that depends on disruption of cadherin/catenin adherens complexes.48–50 Neutrophil recruitment depends on rolling on capillary endothelia, tight binding via intercellular adhesion molecules (ICAMs), diapedesis, and migration. Why is cleavage of E-cadherin observed so early during the onset of acute pancreatitis? The answer is provided by the effects of PMN elastase, which contributes to disease severity, with PMN elastase levels being prognostic.50 During the first hour of hyperstimulation-induced rodent acute pancreatitis, CD45-positive (leukocyte common antigen positive) leukocytes infiltrate the pancreatic parenchyma, and PMN elastase and myeloperoxidase activity are elevated. Furthermore, when neutrophil recruitment is prevented by an absence of essential components of leukocyte integrins essential for ICAM binding, as in the CD18-knockout mouse, milder acute pancreatitis results in reduced pancreatic trypsin activity at 1 and 8 hours. It is concluded, therefore, that neutrophils contribute to early trypsinogen activation, elastase activation, and the severity of pancreatic injury in acute pancreatitis by an unknown mechanism.
The ensuing discussion of the presentation by Dr Lerch indicated that elastase is present in other inflammatory cells such as macrophages; and that breakdown of the intercellular adherens junctions may account for some of the changes of pancreatitis including inhibition of secretion. The discussion raised the possibility that neutrophil activation and/or elastase could perhaps be inhibited by a number of compounds to reduce the severity of pancreatitis. In addition, inhibition of the early production of chemokines by acinar cells themselves, which may account for neutrophil activation in pancreatitis, is a possible strategy for therapeutic intervention. Another point raised is that there is no need for colocalization of lysosomes and ZGs, an early event in pancreatitis, for neutrophil activation to occur; and it remains a challenge to explain why neutrophils cause and/or exacerbate trypsinogen activation.
PHYSIOLOGY SOCIETY LECTURE: ABERRANT CA2+ SIGNALING, BICARBONATE SECRETION, AND PANCREATITIS
Shmuel Muallem (Texas) elaborated the view that abnormal Ca2+ signaling initiates pancreatitis,51 with a variety of factors, including protease-activated receptors (notably protease-activated receptor type 2 [PAR2]) modulating systemic impact.52 Injury occurs via abnormal elevations of [Ca2+]I, maintained by Ca2+ influx into the cell, without which noxious effects are not observed (Fig. 2). Thus, the identity, localization, and regulation of pancreatic acinar cell Ca2+ influx channels, notably CRAC and SOCs, are critical.53–56 Scaffolding proteins regulate Ca2+ entry and Ca2+ release within the cytosol, whereas the proposed SOC may be canonical transient receptor potential channels (TRPC), implicated from several studies to be SOCs and found as TRPC-1, TRPC-3, and TRPC-6 in the pancreatic acinar cell.53 Upon stimulation of the cell and cytosolic Ca2+ release leading to Ca2+ store depletion, TRPC-3 translocates from vesicles to the apical plasma membrane, allowing Ca2+ entry by an IP3– and Homer 1–dependent mechanism.54
FIGURE 2.
Maintenance and effects of toxic cytosolic Ca2+ concentrations ([Ca2+]I) in pancreatic acinar cells. Toxic [Ca2+]I builds up in response to cell insult from bile salts, nonoxidative ethanol metabolites, or FAs, which induce excessive release from intracellular stores and inhibit Ca2+ clearance mechanisms. Premature activation of trypsinogen and other digestive enzymes occurs, with basolateral secretion of some activated enzymes into the interstitial space, contributing to cell injury and pancreatitis. This sequence depends on continued Ca2+ entry into the cell, without which toxicity does not occur. Ductal secretion of HCO3− offers one means of protection, whereby clearance of activated enzymes may be hastened. GPR indicates G protein–linked receptors; CFTR, cystic fibrosis transmembrane conductance regulator; SLC26, solute carrier protein 26, ion transporters that interact to control pancreatic ductal Cl−/HCO3− exchange.
Another recently identified regulator of CRAC and SOC Ca2+ entry is a member of the stromal interaction molecule family (STIM-1). The molecular basis for the function of STIM-1 has been unraveled using a screening assay based on the translocation of a transcription factor, nuclear factor of activated T cells, which translocates from the cytosol to the nucleus in response to a sustained elevation of [Ca2+]I.55 These studies have shown that STIM-1 activates SOC, CRAC, and TRPC channels by a mechanism that relies on ezrinradixin-moesin (ERM)-dependent binding and lysine-rich region–dependent activation of the channel from its position within the ER. Ca2+ depletion leads to translocation and clustering of the proteins at sites of close apposition between the plasma membrane and the ER, and promotes the action of the lysine-rich region of STIM-1 on channel opening. It is clustering of STIM-1 within the ER near the membrane in response to store depletion of Ca2+ that is the critical event for channel opening; subsequently, STIM-1 binds TRPCs and determines their function.56 The Eα-helix-loop-Fα-helix hand domain of STIM-1 residing in the ER senses ER Ca2+ concentrations, and functions to inhibit STIM-1 activity when the stores are filled.55 Both the ERM domain and the lysine-rich tail of STIM-1 are crucial for the activation of endogenous SOC, ICRAC, and TRPC channels.
The focus then turned to the enormous output of bicarbonate secreted by the pancreas,57–60 which both ethanol and bile increase. In exploring the mechanism of bicarbonate secretion including the role of the cystic fibrosis transmembrane conductance regulator (CFTR), an ABC chloride ion transporter, the actions of the solute carrier protein 26 transporter (SLC26A) family were discussed.58–60 This family of epithelial anion transporters have currently taken center stage in determining how epithelial cells transport a variety of ions including Cl−, bicarbonate, I−, oxalate, formate, and sulfate. SLC26A3 and SLC26A6 exchange Cl− and bicarbonate in the pancreatic duct, and CFTR regulates these SLC26A types, although SLC26A transporters modulate the activity of CFTR also. Interestingly, cerulein pancreatitis is significantly more severe in SLC26A6-knockout mice than in wild-type animals. These data led to the suggestion that acute pancreatitis may be ameliorated by stimulation of pancreatic ductal secretion, in part to flush out activated enzymes and toxic mediators (Fig. 2).
In the discussion of Dr Muallem’s presentation, it was revealed that TRPC-1, TRPC-3, and TRPC-6 are situated predominantly in the lateral membrane of the pancreatic acinar cell, but there is 100% overlap of TRPC and IP3R. Ca2+ influx is necessary to sustain secretion because without Ca2+ in the external medium, secretion is sustained for 5 minutes only. In addition, cystic fibrosis requires 90% loss of CFTR activity, although with a less severe loss, pancreatic exocrine insufficiency may result.
NF-κB, CYTOKINES, AND IMMUNE MECHANISMS
NF-κB in Pancreatitis
Ilya Gukovsky (Los Angeles, Calif) reviewed the role of NF-κB, particularly in pancreatitis, which features two peaks of NF-κB activation early in its pathogenesis.61–65 This molecule is activated in canonical (classical) and noncanonical (nonclassical) pathways, with an atypical pathway, for example, via DNA damage, all of which may occur within pancreatic acinar cells. NF-κB is activated by bile salts63 and FAEEs,64,65 and alongside p38 mitogen-activated protein kinase, contribute to cytokine up-regulation.62 In hyperstimulation-induced acute pancreatitis, NF-κB is activated within 5 minutes, although bombesin, carbachol, and JMV180 do not induce this, and JMV180 prevents cerulein-induced activation.61 When NF-κB activation is prevented or reduced, as in knockout of TNF-αR (which activates NF-κB) or interleukin-1 receptor, hyperstimulation-induced pancreatitis is less severe. Similarly, P50 (a subunit of NF-κB)-knockout mice develop less severe pancreatitis. The relative roles of [Ca2+]I elevations, PKC isoforms (δ and ε), PI3K (γ), trypsinogen activation, and reactive oxygen species (ROS) are unclear, but it appears that Ca2+ and PKC-δ and/or -ε may activate NF-κB independently. It remains to be determined what the mechanisms of upstream signaling are in pancreatitis and the role of NF-κB in cell death responses.
Dr Gukovsky’s presentation prompted considerable discussion, initially on the possible pathways, which all depend largely if not entirely on external stimuli affecting the cell. For example, the second peak of NF-κB activation is caused by inflammatory infiltrate. A role for LTC4 was suggested. Further debate continued on the relative roles of enzyme activation versus cytokine release in determining the severity of pancreatitis. Several lines of experimentation were suggested, including experiments with BAPTA, and investigation as to whether adenoviral transfer of P65 into the pancreas induces actual acinar cell injury and elevation of serum amylase. Cytokines are critical to the systemic inflammatory response syndrome and the acute respiratory distress syndrome. It was noted that in cardiomyocyte necrosis, ischemia is a key trigger, then a secondary phase occurs during reperfusion with generation of ROS and elevated [Ca2+]I. The relative contribution of inflammatory mediators in acinar cell injury and pancreatitis requires clarification because it is unclear why in producing inflammatory mediators, acinar cells appeared to be overdosing themselves with what should be a protective response. The group pondered whether the relative roles of NF-κB, cytokines, neutrophils, ROS, and Ca2+ could be determined. There is a paradox in that the volume of the pancreas and death of tissue in necrotizing pancreatitis is out of proportion to the systemic effect, and much greater than expected from the death of a similar volume of other tissues.
Inflammation and the Severity of Pancreatitis
Stephen Pandol explored the effects of the inflammatory response in acinar cell injury and pancreatitis.66–70 The production of cytokines by acinar cells,66 neutrophil infiltration and activation,67 stellate cell activation,68 and vascular permeability changes69 all have an effect on disease severity70 and progression to chronicity.68,70 Antineutrophil serum decreases enzyme activation and necrosis in acute pancreatitis, with an accompanying increase in apoptosis. Nicotinamide adenine dinucleotide phosphate (NADPH) oxidase knockdown has the same effect.67 Using measurements of Western blot analysis, generation of ROS and immunocyto-chemistry, NADPH oxidase activity in pancreatitis was demonstrated to be in neutrophils and not pancreatic acinar cells. Guamerin-derived synthetic peptide is a novel specific inhibitor of neutrophil elastase that also inhibits the production of ROS by neutrophils and has been found to reduce the severity of hyperstimulation-induced pancreatitis. Neutrophils seem to convert apoptosis into necrosis, but neutrophil myeloperoxidase (MPO) seems to have no part in this because MPO knockdown was not found to reduce the severity of pancreatitis.67
The discussion of Dr Pandol’s presentation was focused on the sequence of processes from inflammation signaling to necrosis. The mechanisms connecting 1 process to the next in sequence are not determined. Intercellular adhesion molecule 1 is expressed on hyperstimulated cells, which may account for the targeted action of neutrophils, but how apoptosis is converted into necrosis is not known, although the inflammatory response contributes. It was suggested to be a misfortune that evolution has resulted in PMN leukocytes causing trypsinogen activation.
The Role of Neutrophils in Pancreatitis
Ashok Saluja examined questions concerning the stimuli and actions of neutrophils in pancreatic acinar cell death and pancreatitis.71–75 Less severe acute pancreatitis results from hyperstimulation in ICAM-1–knockout mice compared with wild-type animals.72 Similarly, deficiencies of P-selectin or E-selectin, which bind leukocytes to endothelia, result in fewer neutrophils infiltrating the pancreas and less severe pancreatitis. Neutrophil infiltration begins after trypsinogen activation starts but also causes further trypsinogen activation. There are a number of unsolved puzzles. Does Toll-like receptor 4 (TLR-4), a pattern recognition receptor, have a role? This is certainly suggested from experiments in TLR-4–knockout mice, where MPO activity is reduced in both the pancreas and lungs after the induction of acute pancreatitis by hyperstimulation or l-arginine.75 A similar reduction has been found in mice with knockout of CD14, another pattern recognition receptor and coregulator with TLR-4 of responses to endotoxin. No lipopolysaccharide, however, has been found in the experimental animals anywhere, so what activates TLR-4 remains unknown.
The discussion of Dr Saluja’s presentation addressed whether inflammatory signaling alone is sufficient to account for all the processes that occur in pancreatitis. The answer is not evident. In addition, there was a discussion as to why PAR2 has apparently different roles in the pancreas and elsewhere. That is, in the pancreas, PAR2 seems to be protective, whereas in the lung tissue, it seems to worsen damage.
MITOCHONDRIAL INJURY
Mitochondrial Targets and Mediators of Cell Injury
Michael Duchen (London, United Kingdom) pointed to the role of mitochondria in the life and death of cells as an energy source, replete with synthetic enzymes, and containing maternally inherited DNA necessary for respiration.76–80 Mitochondria accumulate Ca2+, integrating the spatiotemporal patterning of cytosolic Ca2+ signals into an energy supply for the rest of the cell.80 Central to mitochondrial function is maintenance of a mitochondrial membrane potential (ΔψM) of 150 to 180 mV across the inner mitochondrial membrane, negative inside, produced by pumping protons into the intermembrane space. Some ROS generation is normal, and their role/effect on ΔψM is unclear, but there are innumerable mitochondrial targets of injury, for example, complexes from excessive ROS generation.77–79 The part played by uncoupling proteins, including uncoupling protein 2, remains unclear, but in uncoupling, protons leak back across the inner membrane.
Mitochondrial dysfunction is the primary cause of some diseases, whereas mitochondrial damage is the final determinant of irreversible cell injury in others.80 Thus, principal questions are how mitochondrial function is compromised, and what happens when this occurs. Tools to dissect mechanisms include oligomycin, which inhibits mitochondrial ATPase; bongkrekate, which inhibits adenine nucleotide translocase (ANT) and thereby ATP exchange from the cytoplasm for mitochondrial ADP; 2-deoxy-d-glucose, which inhibits glycolysis; and pyruvate, on which mitochondrial ATP generation is entirely dependent.76–80 Mitochondria, however, can also be ATP consumers, an important feature of mitochondrial dysfunction, as demonstrated in studies of cardiac myocytes using the mitochondrial dye TMRM. Thus, with laser light application, the ΔψM is lost, but oligomycin prevents the rigor (and cell shortening) that normally results from mitochondrial membrane depolarization. The combination of Ca2+ loading and oxidative and/or nitrosative stress is particularly damaging (Fig. 3), as shown in studies where dyes are oxidized to a fluorescent product, with A amyloid causing a large irreversible mitochondrial depolarization dependent on Ca2+ entry77,79 and NADPH oxidase activation via PKC.79 The effects of Ca2+ loading and oxidative and/or nitrosative stress can be prevented by exclusion of Ca2+ from the medium, blockage of the Ca2+ uniporter (which transports Ca2+ from the cytosol into the mitochondrial matrix), using Ru360 (μ-oxo) bis (trans-formatotetramine ruthenium), cyclosporine A (CyA, which inhibits the mitochondrial permeability transition pore [MPTP]), or knockout of cyclophilin D (CyP-D) (part of the MPTP).77–79 It can also be prevented by antioxidants,79 including 2,2,6,6-tetramethylpiperidyl-1-oxyl radicals and catalase.
FIGURE 3.
Factors controlling cell death in response to pancreatic acinar cell stress. Sustained Ca2+ release induced by toxins results in mitochondrial impairment, which causes release of cytochrome c, more likely if initiator caspases are activated or ROS generation is excessive. Cytochrome c activates effector caspases, which are inhibited by NF-κB, but in turn inhibit necrosis. With marked mitochondrial impairment, however, ATP production is insufficient to support apoptosis, and necrosis results, hastened through the activation of PARP. Activation of cathepsin B and trypsinogen in lysosomes and ZGs contributes to cell injury and deregulated necrotic cell death. PARP indicates poly ADP-ribose polymerase.
Questions raised in the discussion of Dr Duchen’s presentation led him to comment that the principal determinant of ATPase reversal is its thermodynamic equilibrium, that inhibition of respiration increases proton leak, and that the inhibitor protein IF1 and pharmacological agents prevent ATPase reversal. It was also noted that ROS are more important than Ca2+ in the behavior of the MPTP, that reversibility is maintained in vitro as long as glucose is present in the medium, and the extent to which the MPTP is open or closed determines outcome. In vivo, both glycolytic (cytoplasmic) and aerobic (mitochondrial) metabolism may occur, but Dr Duchen stated that caution is required in the interpretation of experiments based on transformed cell lines because these are heavily dependent on glycolysis; damaged cells may also be.
The Role of the MPTP in Apoptotic and Necrotic Cell Death
Andrew Halestrap (Bristol, United Kingdom) continued the mitochondrial theme by recounting how the inner membrane is impermeable to all but a few ions. Toxins or hypoxia causes the inner mitochondrial membrane to become permeable to all solutes less than 1500 d, and triggered by a high Ca2+ in the matrix, mitochondria swell, leak, and become uncoupled.81–85 Sensitization to such damage occurs as a result of oxidative stress. CyP-D is the main target of CyA, which inhibits the MPTP; in CyP-D–knockout mice, there is impaired formation of the MPTP.85 High Ca2+ overcomes the protection of CyA, sanglifehrin A and Cyp-D knockout.83 The MPTP is inhibited by ADP, and the potency of different nucleotides to inhibit the MPTP is dependent on their affinity as substrates of the ANT.81 Nevertheless, mice with a double ANT knockout in their liver mitochondria still show a CyA-sensitive MPTP, although it is less sensitive to Ca2+ and is not blocked by ADP or activated by carboxyatractyloside (a specific inhibitor of the ANT). There is also an interesting question in that it is unclear how the double ANT knockout can survive—there must presumably be another mechanism of mitochondrial ATP export.85
Using the “hot-dog” technique for measuring MPTP opening, where the amount of [3H]-2-deoxy-d-glucose-6-P in mitochondria is used as an indicator of pore opening, ischemic preconditioning has been shown to inhibit MPTP opening.82 Preconditioning protection develops during the second round of ischemia, by decreasing Ca2+-dependent swelling of mitochondria and attenuating mitochondrial protein oxidation, and may inhibit MPTP indirectly by decreasing oxidative stress. Because necrosis probably results from sustained MPTP opening, whereas apoptosis results from transient MPTP opening, preconditioning reduces the likelihood of necrosis.82,85
Discussion of Dr Halestrap’s presentation centered on the role of ANT4, which he commented was unclear, of CyP-D, which has an enzymic action, and fatty acids (FAs), which require ATP to be used, whereas deleterious effects of FAs may be attributable to direct effects on the MPTP. Fatty acids also may have direct uncoupling effects, which would be easy to test. Dr Halestrap noted that swelling via the MPTP leading to rupture is common to all cells.
Menadione-Induced Oxidative Stress Activates 2 Apoptotic Pathways in Pancreatic Acinar Cells
Oleg Gerasimenko (Liverpool, United Kingdom) presented recent work examining mechanisms of calcium signaling and apoptosis in pancreatic acinar cells subjected to a variety of stresses.86–90 Previous work had shown the oxidative stressor menadione to induce repetitive cytosolic Ca2+ spikes, partial mitochondrial depolarization, cytochrome c release, and apoptosis.86 The induction of apoptosis was found to be dependent on opening of the MPTP, blocked by bongkrekic acid, and resulted in activation of caspases 9 (within 5 minutes) and 3 via the intrinsic pathway. Approximately 10% to 15% of the induction of apoptosis, however, was not prevented by specific inhibition of caspase 9, suggesting an alternative mechanism, investigation of which formed the body of the presentation.89,90 By selective use a range of fluorescent dyes, Ca2+ chelators, agents that destroy acid compartments (eg, lysosomes) and specific inhibitors, investigation showed that menadione also induces extrinsic activation of caspase 8 that is Ca2+ independent but relies on the release of cathepsins D and E from lysosomes, and occurs over a longer time course (30–40 minutes). When pancreatic acinar cells are exposed to menadione and caspase 9 is inhibited, activation and activity of caspase 8 are increased, suggesting a compensatory mechanism if the intrinsic apoptotic pathway fails to be activated under conditions of cell injury.90 This may be particularly helpful to the exocrine pancreas where necrotic cell death is potentially lethal to the individual.89
The discussion of Dr Gerasimenko’s work focused on the possible uses of animals with genetic knockouts of caspase 8, cathepsin B, and cathepsin L to establish their contribution to the death mechanism. Dr Gerasimenko also commented that his findings were independent of changes in ADP or ATP; and that in the continuous presence of menadione, there was recovery of ΔψM, as measured by the TMRM probe. The possibility of arylation and/or thiol group formation by menadione was raised.
CELL DEATH PATHWAYS
Mechanisms and Consequences of Apoptosis
Rosario Rizzuto appraised the role of several molecules in cell death, notably Bcl-2, peroxisome proliferator-activated receptor γ coactivator (PGC)-1α, PKC, Shc/p66, and IP3R,91–95 studying their effects on Ca2+ signaling and the likelihood of preventing or inducing cytosolic and/or mitochondrial Ca2+ overload, a key cell death trigger. After IP3 production and Ca2+ release into the cytosol, mitochondria take up Ca2+ and mitochondrial metabolism is stimulated, a pattern that may be perturbed in various ways, and depends on close ER-mitochondrial and nuclear-mitochondrial interactions.91,92 Ca2+ overload and/or release of caspase cofactors from the mitochondria sensitize the cell to apoptosis. A somewhat rhetorical question was whether Bcl-2 is an ion channel because Bcl-2 reduces Ca2+ uptake into the ER, Ca2+ release from the ER into the cytosol, and Ca2+ uptake into mitochondria. The effects of Bcl-2 on Ca2+ flux, however, are independent of the putative MPTP, and likely to be mediated by direct regulation of the IP3R Ca2+ channel. Key underlying questions concerning Bcl-2, Ca2+, and apoptosis are how mitochondrial Ca2+ uptake is directly modulated by physiological and/or pathological conditions, how this might contribute to apoptosis, what the exact role of Bcl-2 is and whether the anti-apoptotic actions of Bcl-2 are explained by its effects on Ca2+ signaling. The apoptotic agent ceramide was used to dissect out these effects, showing that when cytosolic Ca2+ signals and mitochondrial Ca2+ uptake were reduced by Bcl-2, apoptosis was rendered less likely.
Attention was then turned to the role of PGC-1α in adaptive thermogenics, focused on mitochondrial effects.94 Via adrenergic stimulation and cyclic AMP, PGC-1α induces proteomic changes, including up-regulation of uncoupling proteins NRFs and mtTFA, as well as mitochondrial biogenesis, simultaneously reducing mitochondrial Ca2+ uptake. The importance of PGC-1α is well shown in PGC-1α–knockout mice, which have defective adaptive energy metabolism.
As alluded to earlier (see PKC, Alcohol, and Pancreatitis), PKC isoforms are important players in cellular responses to injury. Classical (α, βI, βII, γ), novel (δ, ε, η, τ), and atypical (ζ, λ) isoforms are variously produced in response to IP3 and diacylglycerol second-messenger production, and inhibit mitochondrial Ca2+ entry (eg, α, β) and/or cytosolic Ca2+ release (eg, α), or increase mitochondrial Ca2+ entry (ζ). Among the actions of PKCs are those on p66Shc (66 kd alternatively spliced isoform of the growth factor adapter Shc that is phosphorylated in response to oxidative stress), which accumulates within mitochondria, inducing increased mitochondrial ROS production that can initiate apoptosis.93,95 Of note, knockout of p66Shc (−/−) extends life span, and the molecule has been proposed to be an integration point for signaling pathways that control longevity and cell death. Work with p66Shc-transfected mouse embryonic fibroblast cells has shown that in response to oxidative stress (H2O2), PKC-β modifies Ca2+ homeostasis via p66Shc, which as a result inhibits mitochondrial Ca2+ import. Despite the reduction in Ca2+ entry, mitochondrial import of p66 Shc, facilitated by propyl isomerase 1–dependent isomerization, increases intramitochondrial ROS production and induces opening of the MPTP.95
The discussion of Dr Rizzuto’s presentation raised questions about transport and localization of p66Shc into the mitochondrial matrix. The only information available is that p66Shc is distributed at a low level in all mitochondria. In addition, these mitochondrial studies provide another example of the sensitivity of ER-mitochondrial interactions because such small changes in ER Ca2+ activity can have profound effects on mitochondrial function and cell fate; and that different PKC isoforms regulate mitochondrial ΔψM in different ways.
Determinants of Death and Survival in Neurons
Pierluigi Nicotera (Leicester, United Kingdom) postulated that the same molecular actors that are responsible for cell death are also responsible for cell survival.96–100 Developing the theme, apoptosis can be seen as directly related to necrosis, as part of the same series of events. Thus, if during apoptosis the PMCA is cleaved or ΔψM and ATP supplies are impaired, necrosis results.96,99,100 Contributors to Ca2+ entry, induced by glutamate accumulation at synaptic or extrasynaptic terminals (which leads to neural excitotoxicity) include N-methyl-d-aspartate receptors, transient receptor potential cation channel, subfamily M, member 7, and acid-sensing ion channels.97,98 Deregulated Ca2+ homeostasis inducing cell death depends on “closed doors.” After N-methyl-d-aspartate receptor, transient receptor potential cation channel, subfamily M, member 7, acid-sensing ion channels, and voltage-gated channels allow excessive Ca2+ entry, calpains are activated, cleaving the Na+/Ca2+ exchanger and PMCA (closing the doors), resulting in prolonged, excessive [Ca2+]I.96,99,100 Differing mitochondrial affinity for Ca2+ and Ca2+-buffering capacity may distinguish signals leading to normal function from those leading to cell death. Thus, in neurons, mitochondria at synaptic clefts have a higher affinity for Ca2+ than extrasynaptic mitochondria and may be more able to cope with Ca2+ signals, although it is not clear whether this results from complex signal differences rather than intrinsic differences between mitochondria in different subcellular locations.
The discussion of Dr Nicotera’s presentation focused on the question of whether there are differences between mitochondria in different subcellular regions, particularly with respect to microdomains of NOS activation. This question remained unanswered. It was hypothesized, however, that differences may be related to the distance of mitochondria from the plasma membrane or to mitochondrial NO handling that occur at the receptor level.
Mechanisms Mediating Cell Death Responses in Pancreatitis
Anna Gukovskaya drew attention to the inverse correlation between apoptosis and necrosis in various models of pancreatitis, depending on species and severity.101–105 In the mouse, high levels of X-linked inhibitor of apoptosis protein are found that inhibit apoptosis, unlike in the rat.105 When pancreatitis is mild, more apoptosis and less necrosis are found, whereas in severe pancreatitis, there is less apoptosis and more necrosis; a distinction can perhaps be drawn between Baccidental necrosis[(mediated by loss of ΔψM and/or loss of ATP production) and Bprogrammed necrosis[(mediated by death receptors and receptor-interacting protein). Because mitochondrial permeabilization is a key cell death mechanism, the mechanisms contributing to permeabilization of pancreatic acinar mitochondria require clarification— hyperstimulation, for example, induces early cytochrome c release and loss of ΔψM.102,103,105 Isolated pancreatic acinar mitochondrial experiments have shown that micromolar concentrations (up to 40 μM) of Ca2+ induce loss of ΔψM (change in TMRM fluorescence) proportional to the Ca2+ concentration, whereas cytochrome c release is maximal at 1 to 2 μM Ca2+. At that Ca2+ concentration, bongkrekic acid or ADP restored ΔψM fully, CyA restored this partially. Reactive oxygen species, however, had no effect. The partial effect of CyA suggested that Ca2+-induced depolarization is not mediated through the classical MPTP in pancreatic acinar mitochondria. Further experiments indicated that for cytochrome c release from pancreatic mitochondria, ΔψM must be maintained, as well as ROS generation—which is negatively regulated by depolarization. Superimposed Ca2+ entry will then induce either apoptosis or necrosis, depending on the severity of the stress to the cell, the loss of ΔψM, and the loss of ATP production.
The discussion of Dr Gukovskaya’s presentation focused on the possibility that the MPTP could contribute to cytochrome c release caused by Ca2+ overload in pancreatic acinar cell mitochondria because bongkrekic acid blocked the Ca2+-induced depolarization, and little mitochondrial swelling was seen on the scattergrams. In addition, the MPTP can open and close without being detectable in this experimental system.
Pancreatic Acinar Cell Necrosis
Robert Sutton (Liverpool, United Kingdom) gave the last presentation on causes, mechanisms, and effects of pancreatic acinar cell necrosis.106–110 Pancreatic necrosis makes a major contribution to the severity of clinical acute pancreatitis, its morbidity, and resulting mortality.107,110 In the face of a lack of biological explanations for premature digestive enzyme activation, acinar cell dysfunction, and cell death for more than a decade, the Ca2+ hypothesis106 has gained more substance and thrown up many important leads.110 Thus, causes of pancreatitis (passage of gallstones through the ampulla of Vater resulting in bile entering the pancreatic duct; ethanol and its metabolites; hyperlipidemia; hyperstimulation; etc) induce excessive Ca2+ release from intracellular stores, that when sustained by continued Ca2+ entry, initiates cell damage and death.108,109 Globalization of elevated [Ca2+]I beyond the perigranular mitochondrial firewall is a consistent feature, at least in part, the result of inhibition of ATP production and/or use. More than 2 decades ago, FAEEs were identified as nonoxidative metabolites of alcohol formed by the combination of ethanol with intracellular FAs through the action of synthase enzymes, present in high concentration within the pancreas, whereas their products, FAEEs, were found in high concentrations within the pancreata of individuals who were acutely intoxicated with ethanol at the time of death. Recent work has shown that FAEEs induce massive Ca2+ release via IP3 receptors from ER Ca2+ stores, sustained by Ca2+ entry and resulting in mitochondrial depolarization, failure of ATP production, and cellular necrosis.109 The FAEEs bind within mitochondria and undergo hydrolysis to release FAs and ethanol, which are subsequently oxidized. In fact, it is the FAs that are mostly responsible for impaired ATP production because inhibition of FAEE hydrolysis prevents mitochondrial impairment, whereas FAs themselves induce sustained global elevations of pancreatic acinar [Ca2+]I, mitochondrial impairment, and cellular necrosis, rescued by patch-pipette delivery of intracellular ATP.110 Thus, among many potential avenues of new prophylactic or therapeutic intervention for pancreatitis, either inhibition of Ca2+ release or inhibition of FAEE hydrolysis could find clinical application.
OVERVIEW OF AREAS FOR FUTURE RESEARCH
Organelle Dysfunction
The link between the initial cellular insult, its immediate effects, and subsequent events in subcellular organelles requires significant further research. Hypotheses should be developed for experimental testing with a high potential for clinical application. As an example, do the effects of secretory block lead to intracellular enzyme activation and/or mitochondrial death cascades? What are the mechanisms connecting these processes? Testable hypotheses relating the combination of high [Ca2+]I and ROS generation to mitochondrial and cell death are very relevant. Many questions exist concerning mitochondria, including whether there is any role for mitochondrial genomic degradation, if there is a minimal MPTP sequence, and what the role of ANT, sensitivity factors (eg, VDAC), or protective factors (eg, PKC) might be. Particularly, productive hypotheses will be those designed to test mechanisms to prevent or reverse deleterious mitochondrial changes.
Disordered Acinar Secretion
Persistently high [Ca2+]I destroys the normal cellular machinery caused in part by decreased cellular ATP, resulting in disassembly of normal cellular functions especially affecting zymogens and their secretion The mechanisms of secretion are only partly understood. There should be significant effort to enhance our understanding of the molecular machinery of secretion, including soluble N-ethylmaleimide sensitive factor attachment receptors (SNARES), to better determine strategies to prevent and/or reverse the secretory blockade of pancreatitis.
Role of Ductal Secretion
Secretion from the pancreatic ductal tree may be important in pancreatitis, to wash away activated enzymes. Ductal secretion may be promoted by modification of ion transporters that are CFTR independent, but the relative importance of intraductal versus intraparenchymal enzyme activity is undetermined. In pancreatitis, disordered secretion includes basolateral exocytosis at the expense of apical secretion, so maintenance of normal acinar cell secretion may be equally important.
New Approaches
As well as new technologies, more in vivo data are required and recognition of the uses and limitations of appropriate models. Alongside the many sophisticated bioscience tools, more human data are needed, both in prophylactic and therapeutic scenarios. Significant impact on human acute pancreatitis may require cocktails of medications, so it is likely that many more clinical trials will be required.
Therapeutic Targets
At present, it would seem that if a specific SOC or CRAC were to be identified on the plasmalemma of the pancreatic acinar cell, Ca2+ entry might be prevented and so injury dependent on an elevated [Ca2+]I could be prevented. It is appreciated that there are many other signaling mechanisms that can mediate injury alone or in combination with Ca2+. There are several windows of opportunity that could alter the natural history of pancreatitis (prevent the disease entirely, modify the early course of severe acute pancreatitis, or prevent progression to chronic disease), which further research along these and other lines offers prospects of exploitation.
ACKNOWLEDGMENTS
The organizers thank the British Society of Gastroenterology (Jean Crabtree, Di Tolfree), the American Gastroenterology Association Foundation for Digestive Health and Nutrition (Don Rockey, Stacey Hinton), the Gastroenterology Research Group (John Del Valle, Vincent W. Yang), the Physiology Society (Prem Kumar, Heidi Adnum), and Solvay Healthcare (Kelly Sherry, Ged Collins) for the support of this workshop.
REFERENCES
- 1.Petersen OH. Ca2+ signalling and Ca2+-activated ion channels in exocrine acinar cells. Cell Calcium. 2005;38:171–200. [DOI] [PubMed] [Google Scholar]
- 2.Menteyne A, Burdakov A, Charpentier G, et al. Generation of specific Ca2+ signals from Ca2+ stores and endocytosis by differential coupling to messengers. Curr Biol. 2006;16:1931–1937. [DOI] [PubMed] [Google Scholar]
- 3.Gerasimenko JV, Sherwood M, Tepikin AV, et al. NAADP, cADPR and IP3 all release Ca2+ from the endoplasmic reticulum and an acidic store in the secretory granule area. J Cell Sci. 2006;119:226–238. [DOI] [PubMed] [Google Scholar]
- 4.Criddle DN, Murphy J, Fistetto G, et al. Fatty acid ethyl esters cause pancreatic calcium toxicity via inositol trisphosphate receptors and loss of ATP synthesis. Gastroenterology. 2006;130:781–793. [DOI] [PubMed] [Google Scholar]
- 5.Petersen OH, Sutton R, Criddle DN. Failure of calcium microdomain generation and pathological consequences. Cell Calcium. 2006;40: 593–600. [DOI] [PubMed] [Google Scholar]
- 6.Parekh AB, Putney JW Jr. Regulation of the calcium release-activated calcium current ICRAC. Physiol Rev. 2005;85:757–810. [DOI] [PubMed] [Google Scholar]
- 7.Parekh AB. Cracking the calcium entry code. Nature. 2006;441: 163–164. [DOI] [PubMed] [Google Scholar]
- 8.Moreau B, Nelson JC, Parekh AB. Biphasic regulation of mitochondrial Ca2+ uptake by cytosolic Ca2+ concentration. Curr Biol. 2006;16:1672–1677. [DOI] [PubMed] [Google Scholar]
- 9.Chang W-C, Parekh AB. Close functional coupling between CRAC channels, arachidonic acid release and leukotriene secretion. J Biol Chem. 2004;279:29994–29999. [DOI] [PubMed] [Google Scholar]
- 10.Chang W-C, Nelson C, Parekh AB. Ca2+ through CRAC channels activates cytosolic phospholipase A2 leukotriene C4 secretion and expression of c-fos through ERK-dependent and independent pathways. FASEB J. 2006;20:2381–2383. [DOI] [PubMed] [Google Scholar]
- 11.Rizzuto R, Pozzan T. Microdomains of intracellular Ca2+: molecular determinants and functional consequences. Physiol Rev. 2006;86:369–408. [DOI] [PubMed] [Google Scholar]
- 12.Pinton P, Leo S, Wieckowski MR, et al. Long-term modulation of mitochondrial Ca2+ signals by protein kinase C isozymes. J Cell Biol. 2004;165:223–232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Chami M, Prandini A, Campanella M, et al. Bcl-2 and Bax exert opposing effects on Ca2+ signalling, which do not depend on their putative pore-forming region. J Biol Chem. 2004;279:54581–54589. [DOI] [PubMed] [Google Scholar]
- 14.Wieckowski MR, Szabadkai G, Wasilewski M, et al. Overexpression of adenine nucleotide translocase reduces Ca2+ signal transmission between the ER and mitochondria. Biochem Biophys Res Commun. 2006;348:393–399. [DOI] [PubMed] [Google Scholar]
- 15.Szabadkai G, Bianchi K, Varnai P, et al. Chaperone mediated coupling of endoplasmic reticulum and mitochondrial Ca2+ channels. J Cell Biol. 2006;175:901–911. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Satoh A, Gukovskaya AS, Nieto JM, et al. PKC-delta and -epsilon regulate NF-kappaB activation induced by cholecystokinin and TNF-alpha in pancreatic acinar cells. Am J Physiol Gastrointest Liver Physiol. 2004;287:G582–G591. [DOI] [PubMed] [Google Scholar]
- 17.Satoh A, Gukovskaya AS, Edderkaoui M, et al. Tumor necrosis factor-alpha mediates pancreatitis responses in acinar cells via protein kinase C and proline-rich tyrosine kinase 2. Gastroenterology. 2005;129:639–651. [DOI] [PubMed] [Google Scholar]
- 18.Gukovskaya AS, Hosseini S, Satoh A, et al. Ethanol differentially regulates NF-kappaB activation in pancreatic acinar cells through calcium and protein kinase C pathways. Am J Physiol Gastrointest Liver Physiol. 2004;286:G204–G213. [DOI] [PubMed] [Google Scholar]
- 19.Satoh A, Gukovskaya AS, Reeve JR Jr, et al. Ethanol sensitizes NF-kappaB activation in pancreatic acinar cells through effects on protein kinase C-epsilon. Am J Physiol Gastrointest Liver Physiol. 2006;291:G432–G438. [DOI] [PubMed] [Google Scholar]
- 20.Lam PP, Cosen Binker LI, Lugea A, et al. Alcohol redirects CCK-mediated apical exocytosis to the acinar basolateral membrane in alcoholic pancreatitis. Traffic. 2007;8:605–617. [DOI] [PubMed] [Google Scholar]
- 21.Pandol SJ, Gukovsky I, Satoh A, et al. Animal and in vitro models of alcoholic pancreatitis: role of cholecystokinin. Pancreas. 2003; 27:297–300. [DOI] [PubMed] [Google Scholar]
- 22.Gukovskaya AS, Mareninova OA, Odinokova IV, et al. Cell death in pancreatitis: effects of alcohol. J Gastroenterol Hepatol. 2006;21 (suppl 3):S10–S13. [DOI] [PubMed] [Google Scholar]
- 23.Gukovsky I, Cheng JH, Nam KJ, et al. Phosphatidylinositide 3-kinase gamma regulates key pathologic responses to cholecystokinin in pancreatic acinar cells. Gastroenterology. 2004;126:554–566. [DOI] [PubMed] [Google Scholar]
- 24.Fisher L, Gukovskaya AS, Youg SH, et al. Phosphatidylinositol 3-kinase regulates Ca2+ signaling in pancreatic acinar cells through inhibition of sarco(endo)plasmic reticulum Ca2+-ATPase. Am J Physiol Gastrointest Liver Physiol. 2004;287:G1200–G1212. [DOI] [PubMed] [Google Scholar]
- 25.Fischer L, Gukovskaya AS, Penninger JM, et al. Phosphatidylinositol 3-kinase facilitates bile acid–induced Ca(2+) responses in pancreatic acinar cells. Am J Physiol Gastrointest Liver Physiol. 2007;292: G875–G886. [DOI] [PubMed] [Google Scholar]
- 26.Chvanov M, Petersen OH, Tepikin A. Free radicals and the pancreatic acinar cells: role in physiology and pathology. Philos Trans R Soc Lond B Biol Sci. 2005;360:2273–2284. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Chvanov M, Gerasimenko OV, Petersen OH, et al. Calcium-dependent release of NO from intracellular S-nitrosothiols. EMBO J. 2006;25:3024–3032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Voronina SG, Barrow SL, Gerasimenko OV, et al. Effects of secretagogues and bile acids on mitochondrial membrane potential of pancreatic acinar cells: comparison of different modes of evaluating DeltaPsim. J Biol Chem. 2004;279:27327–27338. [DOI] [PubMed] [Google Scholar]
- 29.Dolman NJ, Gerasimenko JV, Gerasimenko OV, et al. Stable Golgi-mitochondria complexes and formation of Golgi Ca2+ gradients in pancreatic acinar cells. J Biol Chem. 2005;280:15794–15799. [DOI] [PubMed] [Google Scholar]
- 30.Sherwood MW, Prior IA, Voronina SG, et al. Activation of trypsinogen in large endocytic vacuoles of pancreatic acinar cells. Proc Natl Acad Sci U S A. 2007;104:5674–5679. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Verdugo P. Polymer gel phase transition in condensation-decondensation of secretory products. Adv Poly Sci. 1994;110:145–156. [Google Scholar]
- 32.Nguyen T, Chin W-C, Verdugo P. Role of Ca2+/K+ ion exchange in intracellular storage and release of Ca2+. Nature. 1998;395:908–912. [DOI] [PubMed] [Google Scholar]
- 33.Quesada I, Chin W-C, Jordan J, et al. Mouse mast cell secretory granules can function as intracellular ionic oscillators. Biophys J. 2001;80: 2133–2139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Quesada I, Chin W-C, Verdugo P. ATP-Independent luminal Oscillations and release of Ca2+ and H+ from mast cell secretory granules: implications for signal transduction. Biophys J. 2003;85: 963–970. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Quesada I, Verdugo P. InsP3 signaling induces pulse-modulated Ca2+ signaling in the nucleus of airway epithelial ciliated cells. Biophys J. 2005;88:3946–3953. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Kukor Z, Mayerle J, Kruger B, et al. Presence of cathepsin B in the human pancreatic secretory pathway and its role in trypsinogen activation during hereditary pancreatitis. J Biol Chem. 2002;277: 21389–21396. [DOI] [PubMed] [Google Scholar]
- 37.Szmola R, Kukor Z, Sahin-Tóth M. Human mesotrypsin is a unique digestive protease specialized for the degradation of trypsin inhibitors. J Biol Chem. 2003;278:48580–48589. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Nemoda Z, Sahin-Tóth M. The tetra-aspartate motif in the activation peptide of human cationic trypsinogen is essential for autoactivation control, but not for enteropeptidase recognition. J Biol Chem. 2005; 280:29645–29652. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Nemoda Z, Sahin-Tóth M. Chymotrypsin C (caldecrin) stimulates autoactivation of human cationic trypsinogen. J Biol Chem. 2006;281:11879–11886. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Witt H, Sahin-Tóth M, Landt O, et al. A degradation-sensitive anionic trypsinogen (PRSS2) variant protects against chronic pancreatitis. Nat Genet. 2006;38:668–673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Singh VP, Saluja AK, Bhagat L, et al. Serine protease inhibitor causes F-actin redistribution and inhibition of calcium-mediated secretion in pancreatic acini. Gastroenterology. 2001;120:1818–1827. [DOI] [PubMed] [Google Scholar]
- 42.Singh VP, Saluja AK, Bhagat L, et al. Phosphatidylinositol 3-kinase–dependent activation of trypsinogen modulates the severity of acute pancreatitis. J Clin Invest. 2001;108:1387–1395. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Bhagat L, Singh VP, Song AM, et al. Thermal stress-induced HSP70 mediates protection against intrapancreatic trypsinogen activation and acute pancreatitis in rats. Gastroenterology. 2002;122:156–165. [DOI] [PubMed] [Google Scholar]
- 44.Van Acker GJ, Saluja AK, Bhagat L, et al. Cathepsin B inhibition prevents trypsinogen activation and reduces pancreatitis severity. Am J Physiol Gastrointest Liver Physiol. 2002;283:G794–G800. [DOI] [PubMed] [Google Scholar]
- 45.Saluja AK, Lerch MM, Phillips PA, et al. Why does pancreatic overstimulation cause pancreatitis? Annu Rev Physiol. 2007; 69:249–269. [DOI] [PubMed] [Google Scholar]
- 46.Kruger B, Albrecht E, Lerch MM. The role of intracellular calcium signaling in premature protease activation and the onset of pancreatitis. Am J Pathol. 2000;157:43–50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Halangk W, Lerch MM, Brandt-Nedelev B, et al. Role of cathepsin B in intracellular trypsinogen activation and the onset of acute pancreatitis. J Clin Invest. 2000;106:773–781. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Schnekenburger J, Mayerle J, Kruger B, et al. Protein tyrosine phosphatase kappa and SHP-1 are involved in the regulation of cell-cell contacts at adherens junctions in the exocrine pancreas. Gut. 2005;54:1445–1455. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Vasseur S, Folch-Puy E, Hlouschek V, et al. p8 improves pancreatic response to acute pancreatitis by enhancing the expression of the anti-inflammatory protein pancreatitis-associated protein I. J Biol Chem. 2004;279:7199–7207. [DOI] [PubMed] [Google Scholar]
- 50.Mayerle J, Schnekenburger J, Kruger B, et al. Extracellular cleavage of E-cadherin by leukocyte elastase during acute experimental pancreatitis in rats. Gastroenterology. 2005;129:1251–1267. [DOI] [PubMed] [Google Scholar]
- 51.Kim JY, Kim KH, Lee JA, et al. Transporter-mediated bile acid uptake causes Ca2+-dependent cell death in rat pancreatic acinar cells. Gastroenterology. 2002;122:1941–1953. [DOI] [PubMed] [Google Scholar]
- 52.Namkung W, Han W, Luo X, et al. Protease-activated receptor 2 exerts local protection and mediates some systemic complications in acute pancreatitis. Gastroenterology. 2004;126:1844–1859. [DOI] [PubMed] [Google Scholar]
- 53.Li Q, Luo X, Muallem S. Functional mapping of Ca2+ signaling complexes in plasma membrane microdomains of polarized cells. J Biol Chem. 2004;279:27837–27840. [DOI] [PubMed] [Google Scholar]
- 54.Huang GN, Zeng W, Kim JY, et al. STIM1 carboxyl-terminus activates native SOC, I(crac) and TRPC1 channels. Nat Cell Biol. 2006;8: 1003–1010. [DOI] [PubMed] [Google Scholar]
- 55.Kim JY, Zeng W, Kiselyov K, et al. Homer 1 mediates store- and inositol 1,4,5-trisphosphate receptor–dependent translocation and retrieval of TRPC3 to the plasma membrane. J Biol Chem. 2006; 281:32540–32549. [DOI] [PubMed] [Google Scholar]
- 56.Yuan JP, Zeng W, Huang GN, et al. STIM1 heteromultimerizes TRPC channels to determine their function as store-operated channels. Nat Cell Biol. 2007;9:636–645. [Epub ahead of print]. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Choi JY, Muallem D, Kiselyov K, et al. Aberrant CFTR-dependent HCO3-transport in mutations associated with cystic fibrosis. Nature. 2001;410:94–97. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Ko SB, Zeng W, Dorwart MR, et al. Gating of CFTR by the STAS domain of SLC26 transporters. Nat Cell Biol. 2004;6:343–350. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Wang Y, Soyombo AA, Shcheynikov N, et al. Slc26a6 regulates CFTR activity in vivo to determine pancreatic duct HCO3-secretion: relevance to cystic fibrosis. EMBO J. 2006;25:5049–5057. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Shcheynikov N, Wang Y, Park M, et al. Coupling modes and stoichiometry of Cl-/HCO3-exchange by slc26a3 and slc26a6. J Gen Physiol. 2006;127:511–524. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Gukovsky I, Gukovskaya AS, Blinman T, et al. Early NF-kB activation is associated with hormone-induced pancreatitis. Am J Physiol Gastrointest Liver Physiol. 1998;275:G1402–G1414. [DOI] [PubMed] [Google Scholar]
- 62.Blinman TA, Gukovsky I, Mouria M, et al. Activation of pancreatic acinar cells on isolation from tissue: cytokine upregulation via p38 MAP kinase. Am J Physiol Cell Physiol. 2000;279:C1993–C2003. [DOI] [PubMed] [Google Scholar]
- 63.Vaquero E, Gukovsky I, Zaninovic V, et al. Localized pancreatic NF-kappaB activation and inflammatory response in taurocholate-induced pancreatitis. Am J Physiol Gastrointest Liver Physiol. 2001;280:G1197–G1208. [DOI] [PubMed] [Google Scholar]
- 64.Gukovskaya AS, Mouria M, Gukovsky I, et al. Ethanol metabolism and transcription factor activation in pancreatic acinar cells in rats. Gastroenterology. 2002;122:106–118. [DOI] [PubMed] [Google Scholar]
- 65.Kubisch CH, Gukovsky I, Lugea A, et al. Long-term ethanol consumption alters pancreatic gene expression in rats: a possible connection to pancreatic injury. Pancreas. 2006;33:68–76. [DOI] [PubMed] [Google Scholar]
- 66.Gukovskaya AS, Gukovsky I, Zaninovic V, et al. Pancreatic acinar cells produce, release, and respond to tumor necrosis factor-alpha. Role in regulating cell death and pancreatitis. J Clin Invest. 1997;100:1853–1862. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Gukovskaya AS, Vaquero E, Zaninovic V, et al. Neutrophils and NADPH oxidase mediate intrapancreatic trypsin activation in murine experimental acute pancreatitis. Gastroenterology. 2002; 122:974–984. [DOI] [PubMed] [Google Scholar]
- 68.Lugea A, Nan L, French SW, et al. Pancreas recovery following cerulein-induced pancreatitis is impaired in plasminogen-deficient mice. Gastroenterology. 2006;131:885–899. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Ito Y, Lugea A, Pandol SJ, et al. Substance P mediates cerulein-induced pancreatic microcirculatory dysfunction in mice. Pancreas. 2007;34:138–143. [DOI] [PubMed] [Google Scholar]
- 70.Pandol SJ, Saluja AK, Imrie CW, et al. Acute pancreatitis: bench to the bedside. Gastroenterology. 2007;132:1127–1151. [DOI] [PubMed] [Google Scholar]
- 71.Bhatia M, Wallig MA, Hofbauer B, et al. Induction of apoptosis in pancreatic acinar cells reduces the severity of acute pancreatitis. Biochem Biophys Res Commun. 1998;246:476–483. [DOI] [PubMed] [Google Scholar]
- 72.Frossard JL, Saluja A, Bhagat L, et al. The role of intercellular adhesion molecule 1 and neutrophils in acute pancreatitis and pancreatitis-associated lung injury. Gastroenterology. 1999; 116:694–701. [DOI] [PubMed] [Google Scholar]
- 73.Hietaranta AJ, Singh VP, Bhagat L, et al. Water immersion stress prevents caerulein-induced pancreatic acinar cell NF-kappa B activation by attenuating caerulein-induced intracellular Ca2+ changes. J Biol Chem. 2001;276:18742–18747. [DOI] [PubMed] [Google Scholar]
- 74.Singh VP, Bhagat L, Navina S, et al. PAR-2 protects against pancreatitis by stimulating exocrine secretion. Gut. 2006;56: 958–964. [Epub ahead of print]. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Dawra R, Sharif R, Phillips P, et al. Development of a new mouse model of acute pancreatitis induced by administration of L-arginine. Am J Physiol Gastrointest Liver Physiol. 2007;292:G1009–G1018. [DOI] [PubMed] [Google Scholar]
- 76.Jacobson J, Duchen RR. Mitochondrial oxidative stress and cell death in astrocytesVrequirement for stored Ca2+ and sustained opening of the permeability transition pore. J Cell Sci. 2002;115: 1175–1188. [DOI] [PubMed] [Google Scholar]
- 77.Abramov AY, Canevari L, Duchen MR. Beta-amyloid peptides induce mitochondrial dysfunction and oxidative stress in astrocytes and death of neurons through activation of NADPH oxidase. J Neurosci. 2004;24:565–575. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Hausenloy D, Wynne A, Duchen MR, et al. Transient mitochondrial permeability transition pore opening mediates preconditioning induced protection. Circulation. 2004;109:1714–1717. [DOI] [PubMed] [Google Scholar]
- 79.Abramov AY, Canevari L, Duchen MR. Expression and modulation of an NADPH oxidase in mammalian astrocytes. J Neurosci. 2005; 25:9176–9184. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Davidson SM, Duchen MR. Calcium microdomains and oxidative stress. Cell Calcium. 2006;40:561–574. [DOI] [PubMed] [Google Scholar]
- 81.McStay GP, Clarke SJ, Halestrap AP. Role of critical thiol groups on the matrix surface of the adenine nucleotide translocase in the mechanism of the mitochondrial permeability transition pore. Biochem J. 2002;367:541–548. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Lim KHH, Javadov SA, Das M, et al. The effects of ischaemic preconditioning, diazoxide and 5-hydroxydecanoate on rat heart mitochondrial volume and respiration. J Physiol. 2002;545:961–974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Clarke SJ, McStay GP, Halestrap AP. Sanglifehrin A acts as a potent inhibitor of the mitochondrial permeability transition and reperfusion injury of the heart by binding to cyclophilin-D at a different site from cyclosporin A. J Biol Chem. 2002;277:34793–34799. [DOI] [PubMed] [Google Scholar]
- 84.Das M, Parker JE, Halestrap AP. Matrix volume measurements challenge the existence of diazoxide/glibenclamide-sensitive KATP channels in mitochondria. J Physiol. 2003;547:893–902. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Halestrap AP. Calcium, mitochondria and reperfusion injuryVA pore way to die. Biochem Soc Trans. 2006;34:232–237. [DOI] [PubMed] [Google Scholar]
- 86.Gerasimenko JV, Gerasimenko OV, Palejwala A, et al. Menadione-induced apoptosis: roles of cytosolic Ca2+ elevations and the mitochondrial permeability transition pore. J Cell Sci. 2002;115: 485–497. [DOI] [PubMed] [Google Scholar]
- 87.Gerasimenko JV, Maruyama Y, Yano K, et al. NAADP, mobilizes Ca2+ from a thapsigargin-sensitive store in the nuclear envelope by activating ryanodine receptors. J Cell Biol. 2003;163:271–282. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Gerasimenko JV, Flowerdew SE, Voronina SG, et al. Bile acids induce Ca2+ release from both the ER and acidic intracellular Ca2+ stores through activation of IP3Rs and RyRs. J Biol Chem. 2006;281: 40154–40163. [DOI] [PubMed] [Google Scholar]
- 89.Criddle DN, Gerasimenko JV, Baumgartner HK, et al. Calcium signalling and pancreatic cell death: apoptosis or necrosis? Cell Death Differ. 2007;14:1285–1294. [Epub ahead of print]. [DOI] [PubMed] [Google Scholar]
- 90.Baumgartner HK, Gerasimenko JV, Thorne C, et al. Caspase-8–mediated apoptosis induced by oxidative stress is independent of the intrinsic pathway and dependent on cathepsins. Am J Physiol Gastrointest Liver Physiol. 2007;293:G296–G307. [Epub ahead of print]. [DOI] [PubMed] [Google Scholar]
- 91.Szabadkai G, Simoni AM, Chami M, et al. Drp-1–dependent division of the mitochondrial network blocks intraorganellar Ca2+ wavesand protects against Ca2+-mediated apoptosis. Mol Cell. 2004;16:59–68. [DOI] [PubMed] [Google Scholar]
- 92.Cipriani G, Rapizzi E, Vannacci A, et al. Nuclear poly(ADP-ribose) polymerase-1 rapidly triggers mitochondrial dysfunction. J Biol Chem. 2005;280:17227–17234. [DOI] [PubMed] [Google Scholar]
- 93.Giorgio M, Migliaccio E, Orsini F, et al. Electron transfer between cytochrome c and p66Shc generates reactive oxygen species that trigger mitochondrial apoptosis. Cell. 2005;129:221–233. [DOI] [PubMed] [Google Scholar]
- 94.Bianchi K, Vandecasteele G, Carli C, et al. Regulation of Ca2+ signalling and Ca2+ mediated cell death by the transcriptional coactivator PGC-1. Cell Death Differ. 2006;13:586–596. [DOI] [PubMed] [Google Scholar]
- 95.Pinton P, Rimessi A, Marchi S, et al. Protein kinase C beta and prolyl isomerase 1 regulate mitochondrial effects of the life-span determinant p66Shc. Science. 2007;315:659–663. [DOI] [PubMed] [Google Scholar]
- 96.Schwab BL, Guerini D, Didszun C, et al. Cleavage of plasma membrane calcium pumps by caspases: a link between apoptosis and necrosis. Cell Death Differ. 2002;9:818–831. [DOI] [PubMed] [Google Scholar]
- 97.Chami M, Ferrari D, Nicotera P, et al. Caspase-dependent alterations of Ca2+ signaling in the induction of apoptosis by hepatitis B virus X protein. J Biol Chem. 2003;278:31745–31755. [DOI] [PubMed] [Google Scholar]
- 98.Nicotera P, Bano D. The enemy at the gates. Ca2+ entry through TRPM7 channels and anoxic neuronal death. Cell. 2003;115:768–770. [DOI] [PubMed] [Google Scholar]
- 99.Bano D, Young KW, Guerin CJ, et al. Cleavage of the plasma membrane Na+/Ca2+ exchanger in excitotoxicity. Cell. 2005;120:275–285. [DOI] [PubMed] [Google Scholar]
- 100.Bano D, Munarriz E, Chen HL, et al. The plasma membrane Na+/Ca2+ exchanger is cleaved by distinct protease families in neuronal cell death. Ann N Y Acad Sci. 2007;1099:451–455. [DOI] [PubMed] [Google Scholar]
- 101.Gukovskaya AS, Perkins P, Zaninovic V, et al. Mechanisms of cell death after pancreatic duct obstruction in the opossum and the rat. Gastroenterology. 1996;110:875–884. [DOI] [PubMed] [Google Scholar]
- 102.Gukovskaya AS, Gukovsky I, Jung Y, et al. Cholecystokinin induces caspase activation and mitochondrial dysfunction in pancreatic acinar cells. Roles in cell injury processes of pancreatitis. J Biol Chem. 2002;277:22595–22604. [DOI] [PubMed] [Google Scholar]
- 103.Gukovskaya AS, Pandol SJ. Cell death pathways in pancreatitis and pancreatic cancer. Pancreatology. 2004;4:567–586. [DOI] [PubMed] [Google Scholar]
- 104.Wang YL, Hu R, Lugea A, et al. Ethanol feeding alters death signaling in the pancreas. Pancreas. 2006;32:351–359. [DOI] [PubMed] [Google Scholar]
- 105.Mareninova OA, Sung KF, Hong P, et al. Cell death in pancreatitis: caspases protect from necrotizing pancreatitis. J Biol Chem. 2006;281:3370–3381. [DOI] [PubMed] [Google Scholar]
- 106.Ward JB, Petersen OH, Jenkins SA, et al. Is an elevated concentration of acinar cytosolic free ionised calcium the trigger for acute pancreatitis? Lancet. 1995;346:1016–1019. [DOI] [PubMed] [Google Scholar]
- 107.Neoptolemos JP, Raraty M, Finch M, et al. Acute pancreatitis: the substantial human and financial costs. Gut. 1998;42:886–891. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Raraty M, Ward J, Erdemli G, et al. Calcium-dependent enzyme activation and vacuole formation in the apical granular region of pancreatic acinar cells. Proc Natl Acad Sci U S A. 2000;97: 13126–13131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Criddle DN, Raraty MG, Neoptolemos JP, et al. Ethanol toxicity in pancreatic acinar cells: mediation by nonoxidative fatty acid metabolites. Proc Natl Acad Sci U S A. 2004;101:10738–10743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Petersen OH, Sutton R. Ca2+ signalling and pancreatitis: effects of alcohol, bile and coffee. Trends Pharmacol Sci. 2006;27:113–120. [DOI] [PubMed] [Google Scholar]



