Next Article in Journal
Duodenal Adenocarcinoma: The Relationship between Type of Surgery and Site of Recurrence in a Spanish Cohort
Previous Article in Journal
Prevention of Musculoskeletal Injuries in Gastrointestinal Endoscopists
Previous Article in Special Issue
GLP-1RA Essentials in Gastroenterology: Side Effect Management, Precautions for Endoscopy and Applications for Gastrointestinal Disease Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Pathogenesis of Pancreatitis and the Role of Autophagy

by
Ioannis Tsomidis
1,*,
Argyro Voumvouraki
2 and
Elias Kouroumalis
1
1
Laboratory of Gastroenterology and Hepatology, University of Crete Medical School, 71500 Heraklion, Crete, Greece
2
1st Department of Internal Medicine, AHEPA University Hospital, 54621 Thessaloniki, Greece
*
Author to whom correspondence should be addressed.
Gastroenterol. Insights 2024, 15(2), 303-341; https://doi.org/10.3390/gastroent15020022
Submission received: 9 March 2024 / Revised: 31 March 2024 / Accepted: 9 April 2024 / Published: 22 April 2024
(This article belongs to the Special Issue Recent Advances in the Management of Gastrointestinal Disorders)

Abstract

:
The pathogenesis of acute and chronic pancreatitis has recently evolved as new findings demonstrate a complex mechanism operating through various pathways. In this review, the current evidence indicating that several mechanisms act in concert to induce and perpetuate pancreatitis were presented. As autophagy is now considered a fundamental mechanism in the pathophysiology of both acute and chronic pancreatitis, the fundamentals of the autophagy pathway were discussed to allow for a better understanding of the pathophysiological mechanisms of pancreatitis. The various aspects of pathogenesis, including trypsinogen activation, ER stress and mitochondrial dysfunction, the implications of inflammation, and macrophage involvement in innate immunity, as well as the significance of pancreatic stellate cells in the development of fibrosis, were also analyzed. Recent findings on exosomes and the miRNA regulatory role were also presented. Finally, the role of autophagy in the protection and aggravation of pancreatitis and possible therapeutic implications were reviewed.

1. Introduction

The main etiologies of both acute (AP) and chronic pancreatitis (CP) are still gallstones and prolonged alcohol consumption. Their incidence is closely related to the prevalence of gallstone disease and alcohol abuse [1]. Alcohol abuse as a cause of pancreatitis requires consumption of ≥4–5 drinks per day over >5 years [2].
An additional risk factor is smoking, as several studies support an important role for either smoking alone or in combination with alcohol abuse [3,4,5]. The role of hypertriglyceridemia has also been established, as this is the third leading cause of acute pancreatitis [6,7]. It has been reported that approximately 15–20% of individuals with triglyceride levels over 1000 mg/dL will develop acute pancreatitis [8]. Other common etiologies of acute pancreatitis include complications of endoscopic retrograde cholangiopancreatography and autoimmunity, while in a small proportion of cases, no obvious factors can be identified and the term idiopathic pancreatitis is used [1,9]. Diabetes has been associated with an increased risk for pancreatitis [10,11,12]. Finally, medication-related pancreatitis is not a common cause, accounting for fewer than 5% of cases [13]. Drugs strongly associated with acute pancreatitis include azathioprine, 6-mercaptopurine, didanosine, valproic acid, angiotensin-converting enzyme inhibitors, and mesalamine [13].
A distinct group are patients where the first episodes of acute pancreatitis appear before the age of 35. Genetic abnormalities are found in nearly half of them [14]. Interestingly, a genetic mutation in claudin 2 may synergize with alcohol consumption in the development of pancreatitis, indicating the interplay between external and genetic factors in the pathogenesis of pancreatitis [15,16].
Chronic pancreatitis develops in about 10% of patients after the first episode of AP and in about 30% of patients with recurrent AP. Male sex and alcohol abuse are significant risk factors for the transition from AP to CP [17]. Chronic inflammation of the pancreas is caused by acinar and ductular cell injury driven by alcohol, smoking, hypercalcemia, genetic factors, or any combination of the above. It is distinguished from autoimmune CP, which responds to steroid treatment, obstructive CP, and infectious CP. Classic CP can be dominated either by fibrosis or by atrophy [18,19,20]. The relationship of ethanol and smoking to CP has been unequivocally proved. The amount and duration of ethanol consumption have been estimated to be either a median of 5.1 drinks/d or a consumption of up to 110–277 g ethanol/d over 5–25 years, similarly to AP [21,22,23,24,25]. As in AP, a synergy among genetic, immune, and environmental factors may be important [26,27,28,29,30]. Interestingly, post-mortem pathological studies of chronic alcoholics without a previous AP event in their medical history showed the presence of fibrosis and/or ductal calcifications in 47–68% of cases, indicating extensive CP that escaped undiagnosed [25,31,32].
The pathogenesis of both AP and CP is a complex process with many points that have not been fully investigated. Several mechanisms have been incriminated, such as acinar cell auto-digestion, mitochondrial abnormalities, and the involvement of immunity and inflammation. Recently, the role of autophagy has been recognized but not yet fully investigated. Therefore, in this review, we report an outline of the pathophysiology of pancreatitis, with a detailed presentation of the role of autophagy.

2. A Brief Overview of Autophagy

Autophagy is a degradation pathway that allows for the disposition of intracellular waste material including damaged organelles or intracellular pathogens. After lysosomal degradation, most of the final products can be recycled and re-used, supporting the energy system of the cell.
Τhe term autophagy (a Greek word for self-eating) was introduced by Anselmier [33]. The modern concept of autophagy, however, started with the pioneer work of Christian René de Duve in the 1950s, when acid-phosphatase-positive granules were identified in the rat liver [34] and the term lysosome appeared for the first time [35]. The next important step came when the group of Oshumi described a series of fifteen autophagy-related genes (Atgs) involved in Saccharomyces cerevisiae autophagy [36]. Today, more than 40 Atgs have been identified [37]. The importance of autophagy led to the awarding of two Nobel Prizes for Physiology or Medicine, the first to Christian De Duve in 1974 and the second to Yoshinori Ohsumi in 2016 [38]. Historical landmarks of autophagy have been described by Ohsumi [39].
There are several stages in the pathway of autophagy. The initiation stage is followed by the elongation of the phagophore and the autophagosome formation, followed by fusion with lysosomes and degradation of cellular organelles, proteins, and lipids.
Initiation stage and phagophore formation. Autophagy inducers are responsible for the initiation stage. In reality, autophagy is a series of phosphorylations and de-phosphorylations [40]. Three kinases are the main regulators of autophagy, namely the mammalian target of rapamycin (mTOR), the Unc-51 like autophagy activating kinase (ULK1), and the AMP-dependent protein kinase (AMPK) [41]. Autophagy inducers, such as starvation and increased levels of reactive oxygen species (ROS), repress mTOR and activate AMPK, which results in ULK1 activation [42]. Phosphorylation of ULK1 by mTOR reduces its activity, thus decreasing autophagy, while phosphorylation by the AMPK at a different site activates ULK1 and autophagy [43]. Upon induction of autophagy, the ULK1 complex is formed from the assembly of the ULK1, ATG13, FIP200, and ATG101 proteins [44,45]. Autophagy is also upregulated by p38 through inhibition of mTOR, while c-Jun N-terminal kinase1 (JNK1) and BNIP3 (BCL2/adenovirus E1B 19 kDa protein-interacting protein 3) disrupt the B-cell lymphoma 2 (Bcl-2)–Beclin1 complex, inducing autophagy [46,47]. Free Beclin1 binds to Vps34-Vps15 to increase autophagy through the formation of the class III PI3K kinase complex (PI3KC3), consisting of Vps34-Vps15-Beclin1 [48,49]. These complexes lead to the formation of the autophagosome. Then, Vps34 produces phosphatidylinositol-3-phosphate to recruit the effector protein DFCP1, which promotes the development of the double membrane phagophore [50]. The phagophore is generated from the endoplasmic reticulum (ER) or Golgi membranes, mitochondria, and plasma membrane via endocytosis mediated by clathrin.
There are some additional points in the regulation of the initial stages of autophagy. AMPK can negatively regulate mTORC1, either directly through the phosphorylation of mTORC1 activity or indirectly by activating TSC2, which is a strong inhibitor of mTORC1 [51]. Recently, an additional mechanism for mTORC1 activation under energy-rich conditions was described. mTORC1 phosphorylates the protein Pacer, causing the disruption of the complex formed by the proteins Pacer, Syntaxin17 (Stx17), and the homotypic fusion and vacuole protein sorting (HOPS), thus inhibiting the autophagosome maturation mediated by this complex [52,53].
Two additional autophagy regulators have been described. The long non-coding RNA (lncRNA) NBR2 inhibits Beclin 1-dependent autophagy and attenuates the autophagy-induced cell proliferation [54], while Forkhead box O3 (FOXO3), a member of the FOXO subfamily of transcription factors, upregulates autophagy, acting on ULK1, Beclin-1, and LC3 [55].
Expansion (elongation). ULK1 phosphorylation leads to autophagosome formation. The critical step is the phosphorylation of ATG13, leading to the formation of the complex ATG5-ATG12, ATG16L1 [56,57]. This complex and the phosphatidylethanolamine (PE)-LC3 systems are critical for the elongation of phagophores [50]. Pro-LC3 cleaved by ATG4B leads to the generation of the cytosolic form of LC3 (LC3-I). Then, ATG7 processes LC3-I and ATG3 to be conjugated to PE and form LC3-II. The transformation of phagophores into autophagosomes requires the ATG12–ATG5–ATG16 complex and the PE-conjugated LC3II (ATG8) system. Autophagosomes contain materials or cellular organelles destined for degradation.
Autophagosome fusion to the lysosome. This is the final stage of autophagy that allows for the autophagy flux. The autophagosome does not contain hydrolases and the pH is neutral. Fusion with lysosomes forms autolysosomes during the so-called autophagic flux. Overproduction of autophagosomes faster than the flux rate or when flux is repressed will increase the levels of LC3 and p62 [58]. Components that are destined to lysosomal degradation are either labeled by ubiquitin or attached to receptors, such as sequestosome 1 (SQSTM1, also known as p62), and CALCOCO2 (calcium binding and coiled-coil domain 2). These receptors interact with LC3 to deliver the component into the autophagosomes [59,60].
A critical point for fusion to occur is the presence of soluble N-ethylmaleimide-sensitive factor attachment protein receptor (SNARE) proteins localized in opposing membranes of the particles to be fused. Two SNARE complexes mediate the fusion of autophagosomes with lysosomes. The first consists of STX17-SNAP29-VAMP8 [61], and the second is composed of YKT6-SNAP29-STX7 [62]. For a complete fusion, additional proteins are also recruited, such as the HOPS complex, baculovirus IAP repeat containing ubiquitin-conjugating enzyme (BRUCE), and GRASP55, that bind to proteins of the lysosomal membranes, such as Rab-7 and Monensin sensitivity protein 1—Caffeine, calcium, and zinc 1 complexes (Mon1-Ccz1) [63]. ATG8 proteins also contribute to fusion protein recruitment, but they must be removed before final fusion [64]. Rab7 binds to FYCO1 (FYVE and coiled-coil domain-containing 1), ORP1L (oxysterol-binding protein-related protein 1L), and RILP (Rab-interacting lysosomal protein). In the next step, SM (Sec1/Munc-18) family proteins facilitate SNARE complex assembly and zippering. The zippering of these domains fuses the membranes, and SNAREs are now located on the same membrane [65]. The autophagosome–lysosome fusion process requires SNARE complex disassembly on post-fusion membranes [66].
Autophagosome degradation and recycling. After fusion, hydrolases that are active at acidic pH digest the different constituents. In lysosomes, the vacuolar ATPase (vATPase) regulates the import of hydrogen ions to maintain the acidic pH. The same vATPase also induces the transcription factor EB (TFEB). TFEB is phosphorylated in starvation conditions, translocates to the nucleus, and induces the transcription of genes that promote autophagy, including LC3 and p62 [67,68]. mTOR activation decreases TFEB activity, and the autophagic machinery is repressed [69]. TFEB is also a controller of lysosomal biogenesis genes [70,71]. TFEB and ZKSCAN3 are major antagonistic factors during autophagy. ZKSCAN3 is the major transcriptional repressor of autophagy by targeting biogenesis and fusion of autophagosomes and lysosomes [72,73]. TFEB is also a controller of inflammation. Reduction of TFEB leads to exacerbation of inflammation [74]. After degradation, the breakdown products are moved back into the cytosol by lysosomal transporters and re-utilized by the cell [75].
Non-canonical forms of autophagy were reported as leading to similar fusion [76]. Several members of the Atg machinery are not used. Rab9-mediated autophagy functions in cells with atg5 and atg7 deletion. This non-canonical autophagy does not require ATG8/LC3 but is directly regulated by ULK1 [76]. By contrast, LAP is another form of non-canonical autophagy that does not require ULK1/2 but requires ATG8/LC3 conjugation instead and involves ATG5 and ATG7. LAP recruits LC3-II to the phagosomal membrane [77,78,79] and is taken up by macrophages through innate immune receptors, such as Toll-like receptors. In contrast to classical autophagy, the LAPasome is a single membrane vacuole. The term CASM (conjugation of ATG8 to single membranes) was introduced to describe these related pathways [80,81,82]. Detailed descriptions of the autophagy pathway have recently been published [66,83].
Figure 1 summarizes in a simplified diagram the various steps of autophagy.
The main regulators of autophagy are three kinases, namely the mammalian target of rapamycin (mTOR), the Unc-51 like autophagy activating kinase (ULK1), and the AMP-dependent protein kinase (AMPK). ULK1 and AMPK activation promote autophagy, while mTOR activation inhibits autophagy. Beclin 1 and Bcl-2 are important elements in the process. Activation of the TNFR1 leads to caspase 8 activation that cleaves Beclin 1, and the C-terminal fragment inhibits autophagy. The cleavage of Atg4D by caspase-3 generates a fragment, which increases autophagy. The effect of P53 on autophagy depends on localization. Cytoplasmic P53 inhibits autophagy, while nuclear P53 activates AMPK, increasing autophagy.

2.1. Mitophagy

A specialized form of autophagy that is pertinent in pancreatitis is mitophagy. It selectively degrades damaged mitochondria irrespective of the cause of damage [84,85]. Mitophagy is induced by two signal pathways, the PINK1 (PTEN-induced putative kinase 1)-PARKIN (parkin RBR E3 ubiquitin protein ligase) pathway and the PINK1/PARKIN-independent pathway [86,87]. PINK1 is aggregated into the inner mitochondrial membrane (IMM) in normal mitochondria through the activity of the TOM (translocase of the outer mitochondrial membrane) and TIM23 (translocase of inner mitochondrial membrane 23) proteins. PINK1 is cleaved by PARL (presenilin-associated rhomboid like). During severe oxidative stress, impaired mitochondria are not capable of PINK1 seggregation into the IMM. PINK1 associates with TOM and accumulates on the outer mitochondrial membrane (OMM) [88,89], where it recruits and activates Parkin from the cytoplasm [90]. Parkin ubiquitinates several OMM proteins, such as mitofusin 1 and mitofusin 2, voltage-dependent anion channel (VDAC), and Miro [91,92]. Cargo receptor proteins, such as p62, OPTN (optineurin), and CALCOCO2, bind to these OMM proteins to start autophagosome formation [91,93] and subsequent fusion with lysosomes. Mitophagy may be upregulated by phosphorylation of OPTN via the activation of TBK1 (TANK-binding kinase 1) [94,95].
PINK1/Parkin-independent mitophagy requires interaction of LC3II through the LC3-interacting region (LIR) with OMM proteins, such as FUN14 domain containing 1 (FUNDC1), NIP3-like protein X (Nix/Bnip3L), and Bcl-2/adenovirus E1B (Bnip3). The interaction leads damaged mitochondria to the autophagosomes and lysosomal fusion [86].
Details on mitophagy mechanisms have recently been published [96,97,98,99,100].
There are several other specialized forms of autophagy, but their role in pancreatitis has not been investigated [86].

2.2. Autophagy and Immunity

Autophagy is implicated in the regulation of the immune system [101,102], particularly in the regulation of innate immunity in macrophages [103,104]. Interestingly, there is evidence that high autophagic activity is implicated in acquired immunity as well because it maintains the differentiation and function of regulatory T (Treg)-cells [105] and γ δ T-cells [106].

2.3. Autophagy and Cell Death

Autophagy is mostly a protective cellular mechanism supporting cell survival. However, it may turn into a cellular death mechanism through its effect on apoptosis [107]. Autophagy is tightly related to apoptosis. These two pathways are affecting each other, being mutually exclusive [108]. Autophagy reduces the induction of caspase-dependent apoptosis, and apoptosis-associated caspase activation suppresses the autophagic process. Yet, autophagy may induce apoptosis or necrosis, while autophagy itself may degrade the cytoplasm, leading to autophagic cell death [109,110]. The balance between p53 and AKT/mTOR is crucial for the fate of cells [111]. Autophagy also induces a newly described mechanism of cell death named ferroptosis [112,113]. Many proteins vital for autophagy (like ATGs) also participate in ferroptosis. Additionally, activators of ferroptosis, such as erastin, initiate autophagosome formation, while activation of autophagy led to ferroptotic death, possibly by increasing ferrous availability through ferritinophagy [114,115].

2.4. Autophagy and Inflammation

Autophagy is also implicated in the inflammatory response. Inflammasomes and autophagy affect each other. The same inhibitory mechanisms are involved, but they are regulated by different pathways. Autophagy could either repress the assembly of the nucleotide-binding oligomerization domain-like receptor family pyrin domain containing 3 (NLRP3) inflammasome [116] or eliminate active inflammasomes, particularly in macrophages [117]. In addition, the degradation of damaged organelles by autophagy does not allow for the production of more danger-associated molecular patterns (DAMPS) that would further activate NLRP3 inflammasomes [118]. DAMPS activation of NLRP3 leads to pyroptosis through the activation of procaspase 1 activation followed by the production of IL-1β and IL-18 [119]. This negative interplay between autophagy and inflammasome can become positive. Autophagy may induce NLRP3 activation by initiating NF-kB nuclear translocation, leading to pyroptosis [120,121].

3. Pathogenesis of Acute Pancreatitis

Various pathways have been implicated in the complex pathogenesis of AP. Thus, pathological calcium signal transduction, mitochondrial dysfunction, premature activation of trypsinogen in acinar cells and macrophages, endoplasmic reticulum stress (ERs), unfolded protein reaction (UPR), and autophagy impairment have been investigated mostly in animal models [122,123]. Earlier experimental evidence indicated that both acinar cells and ductal cells participate in the pathogenesis of acute pancreatitis. Recently, exosomes that contain proteins, nucleic acids, and lipids have been incriminated in the evolution of AP [124,125].
Abnormalities of intracellular organelles of acinar cells are the basis of the pathogenesis of acute pancreatitis in close association with abnormalities of water and electrolyte secretion by the ductular cells [126,127]. Pancreatic enzyme secretion is blocked in edematous and necrotizing pancreatitis models [128]. On the other hand, pancreatic fluid secretion is four- to five-fold increased at the early stages of pancreatitis, indicating that a defense mechanism is activated to attenuate the severity of the disease [129]. Secretin administration reduces the severity of cerulein-induced pancreatitis [130], a fact that supports the protective effect of this ductal hypersecretion, although this has been disputed [131,132]. The interplay between ductular and acinar cells is also supported by studies demonstrating that pancreatic duct obstruction alone can modify acinar cell membrane trafficking and the evolution of pancreatitis [133]. This is possibly due to the increased intraductal pressure, exposure of cells to bile acids, and acidification of the lumen [134,135,136,137]. An increased intraductal pressure can activate the mechanoreceptor PIEZO1 in the acinar cells to trigger the abnormal calcium signaling discussed below [135] in concert with inflammation and activation of the signal transducer and activator of transcription 3 (STAT3) pathway [138]. Acidification of the pancreatic lumen activates the transient receptor potential vanilloid 1 (TRPV1) in the sensory neurons and causes acute pancreatitis [134]. PIEZO1-mediated and TRPV1-mediated mechanisms of AP are considered to be the main underlying mechanisms for post-ERCP and gallstone pancreatitis [135,139]. Bile acids, on the other hand, can cause mitochondrial dysfunction and damage of the ductal cells [126] that exposes acinar cells to high bile acid concentrations, leading to their death [136,140,141].
These pathogenetic mechanisms are further analyzed.

3.1. Cellular Mechanisms Involved in AP Pathogenesis

3.1.1. Ca++ Signaling and Mitochondrial Dysfunction

Under normal conditions, Ca++ is released from the ER in association with zymogen exocytosis and production of ATP in the mitochondria [142]. This is only a transient increase in cytosolic Ca++, as two ATP-driven calcium channels rapidly reduce the cytosolic calcium. The smooth ER Ca++ channels (SERCAs) send Ca++ back into the ER, while the plasma membrane Ca++ channels (PMCAs) transport Ca++ out of the cell [142]. Protracted elevation of Ca++ concentration in acinar cells initiates activation of pro-inflammatory pathways, such as premature trypsinogen activation, activation of the nuclear factor kappa-light-chain-enhancer of activated B cells (NF-kB), and mitochondrial dysfunction leading to cell death [143,144,145]. Alcohol and bile acids can cause a sustained pathological cytosolic calcium elevation through the inositol 1,4,5-trisphosphate receptor (Ins (1,4,5) P3R) signaling pathway. Alcohol metabolites in the acinar cells open Ins (1,4,5) P3Rs, which are Ca++ channels located in the ER [146,147], thus increasing Ca++ release from the ER lumen [145,148,149]. Increased Ca++ concentration activates the calcium channel protein 1 (ORAI1) to further increase Ca entry into the cell from the outside, maintaining the lethal cellular calcium concentration [150,151]. On the other hand, ductal obstruction, as observed in post-ERCP and gallstone pancreatitis, increases Ca++ entry from the outside through the mechanoreceptor PIEZO1, as mentioned before [135]. Moreover, the high calcium concentration opens the mitochondrial permeability transition pores (MPTP), abolishing the membrane potential needed to generate ATP [145,152,153]. In turn, ATP depletion completes a vicious circle, maintaining the high Ca++ concentration by disrupting the ATP-dependent SERCAs and PMCAs’ clearance of excessive calcium. ATP depletion also impairs other ATP-driven protective mechanisms, such as autophagy and the UPR [122,142], promoting, therefore, acinar cell necrosis.

3.1.2. Mitochondrial Dysfunction

Mitochondrial abnormalities of acinar cells are found in all forms and models of pancreatitis. They cause reduction of ATP synthesis, increased ROS production, and impairment of calcium transport [122,152,154,155,156]. In pancreatitis, there is permeabilization of the mitochondrial membrane due to a sustained opening of the MPTPs, the nonspecific channels crossing both IMM and OMM [157,158,159]. Opening of the MPTPs allows for the uncontrolled entry of water and solutes less than 1500 Da into the matrix, leading to inhibition of ATP synthesis and cellular necrosis. On the other hand, inhibition of the MPTP opening attenuates ATP depletion and acute experimental AP [152]. Not only mitochondrial Ca++ overload, but also increased reactive oxygen species (ROS) generation, cause MPTP opening. Peptidylprolyl isomerase D (cyclophilin D-CypD) is an essential mitochondrial protein around which the MPTPs are organized. CypD inhibition for any reason will block MPTPs’ opening and prevent AP [122,154]. Individual mitochondria form a network that is involved in the mitochondrial activity [160]. In addition, abnormal mitochondrial membrane function releases mitochondrial contents into cytosol, including cytochrome c, which lead to cell death [161].
Aerobic metabolism in mitochondria generates most of the ATP, and only a small quantity is produced by glycolysis. In pancreatitis, anaerobic conditions predominate due to microvascular abnormalities and relative hypoxia of pancreatic tissues. Therefore, the generation of ATP is reduced and not sufficiently replaced by anaerobic glycolysis [30]. Acinar enzyme secretion and ductular bicarbonate production are also significantly reduced [140,162,163]. It should be noted that bile and fatty acids also inhibit ATP production from both sources in acinar and ductal cells [140,164].
An additional difficulty in the pathophysiology of pancreatitis is that current evidence indicates the formation of an interconnected system by different organelles in the acinar cell. Damage of one organelle can lead to failure of the entire network. Thus, the multiple organelle abnormalities found in acute pancreatitis resemble the chicken and egg problem, as it is difficult to dissect the different components responsible for the induction and evolution of AP [122,165]. This is best exemplified in the involvement of the endoplasmic reticulum in the pathophysiology of AP.

3.1.3. Endoplasmic Reticulum (ER) Stress

Mitochondria and ER are closely associated with membrane domains [166,167] providing the amount of calcium required for ATP generation [168,169]. Disruption of the ER–mitochondria interconnections leads to pathologic Ca++ signaling in the acinar cell and low ATP levels [168].
ER stress is the excessive accumulation of misfolded or unfolded proteins within the ER lumen observed when the capacity of the ER to eliminate these proteins is overwhelmed [170]. The pancreas is prone to ER stress because acinar cells produce a large quantity of proteins daily, such as trypsinogen, lipase, and several other lysosomal enzymes [171,172]. ER stress is found frequently in AP and can be triggered by hypoxia, alcohol consumption, Ca++ overload, and oxidative stress [173]. ER over-activation may be an important mechanism that initiates and exacerbates pancreatic injury [174]. During ER stress, acinar cells activate the UPR, which is a strictly controlled signaling pathway that blocks protein translation and synthesis. Meanwhile, the UPR also increases protein folding and the degradation of misfolded proteins, both of which relieve ER stress. UPR uses three functional pathways, namely the inositol-requiring enzyme 1 (IRE1), the activating transcription factor 6 (ATF6), and the protein kinase RNA-like ER kinase (PERK) pathways [175,176,177]. The downstream signals of the IRE1 and ATF6 pathways activate the transcription factors cATF6 and spliced X-box binding protein 1 (sXBP1). These transcription factors increase the synthesis of factors used for ER expansion and chaperones for protein folding [178,179]. They also induce autophagy to recycle misfolded proteins [180]. When UPR is overwhelmed, the apoptotic pathway is activated. On the other hand, the PERK pathway is the terminal response, where its downstream transcriptional factor CEBP homologous protein (CHOP) initiates apoptosis and inflammation [177,181,182,183]. UPR also activates the NF-κB inflammatory pathway, leading to exacerbation of acinar cell inflammation and cell necrosis. Therefore, NF-kB inhibitors, such as IL-10, can block ER stress, reduce pro-inflammatory cytokines, such as TNF-α, IL-1, and IL-6, and delay pancreatic inflammation [184]. CHOP can also induce autophagy, but, in the end, it promotes cell death during prolonged ER stress. Interestingly, 3-hydroxy-3-methylglutaryl-CoA (HMG-CoA) inhibitors, which are used in clinical practice, promote the UPR and may be used to prevent the recurrence of AP [185,186].

3.1.4. Trypsinogen Activation

Observations of pancreatic autolysis in postmortem studies made by the Austrian pathologist Chiari in 1896 [187] formed the basis of the long-held trypsin-centered theory of pancreatic injury.
Trypsinogen activation is the most widely studied pathogenetic mechanism of AP. Premature trypsinogen activation is inhibited by the presence of trypsin inhibitors and zymogen granule release [188]. Alcohol and bile acids stimulate the synthesis of lysosomal digestive enzymes and inhibit the release of zymogen granules at the apex of acinar cells. The lysosome and zymogen granules fuse with one another in the so-called co-localization process [165,188]. Lysosomal cathepsin B, in turn, activates trypsinogen, and both trypsin and cathepsin B are liberated [189]. Cathepsin B release leads to necroptosis, a regulated form of necrosis [190,191] that is mediated by the receptor-interacting protein kinases 1-3 (RIP1-RIP3) and the mixed lineage kinase domain-like (MLKL) pathway [192,193]. MLKL is phosphorylated and oligomerized by RIP3, and the oligomeres are translocated into the plasma membrane, where they cause membrane puncture and spillage of cellular contents [194]. Inhibition of the RIP1–RIP3 by the inhibitor of RIP1 necrostatin attenuates acinar cell injury and can be used as AP therapy [190,191,194]. Furthermore, GSK2982772, a novel RIP1 inhibitor, represses necroptosis and inflammation [195,196] and may be tested in AP treatment [194]. In addition, lysosomal membrane disruption activates caspase 3, which initiates apoptosis through mitochondrial release of cytochrome c [192,193].
Currently, premature trypsinogen activation in acinar cells is considered the central mechanism in the pathogenesis of AP [197]. Recent findings, however, indicate a more complex problem as trypsinogen activation was also observed in macrophages [198,199], demanding further investigations. Macrophage activation of trypsinogen induced translocation of NF-kB and the production of inflammatory cytokines. Cathepsin B-knockout mice without trypsinogen activation in macrophages developed less severe pancreatitis compared to controls [198]. Moreover, another protease, cathepsin D, is expressed in pancreatic acinar cells and macrophages regulating disease severity by activating cathepsin B. Its effect is minimal in the early phase of pancreatitis and much greater in the later, inflammatory cell phase of the disease [200]. These findings challenge the long-held notion that premature trypsinogen activation occurs exclusively within the acinar cells.
Finally, the strongest support for the trypsin-centered theory comes from the identification of mutations in the trypsinogen gene PRSS1 in hereditary pancreatitis, an uncommon form of pancreatitis with autosomal-dominant inheritance [201].
Trypsin activation affects bicarbonate secretion in ductular cells as well. This was attributed to the activation of the basolateral protease activated receptor-2 (PAR-2) [202]. However, luminal administration of trypsin or PAR-2 activating peptide repressed bicarbonate production [203]. Similarly, the severity of experimental pancreatitis can be either reduced or increased after activation of PAR-2 [204,205,206]. Therefore, the matter requires further investigation. Mechanisms of acinar cell injury are summarized in Figure 2.
ETOH and bile acid release cause acute pancreatitis through three mechanisms. (1) They increase Ca++ release from the ER via the InsP3R pathway. Ca++ overload increases the permeability of MPTP, leading to ATP depletion, which blocks SERCA and PMCA and sustains Ca++ overload. Increased Ca++ overload activates trypsinogen and inflammatory signaling pathways but also causes mitochondrial dysfunction, leading to apoptosis and necrosis. (2) They inhibit the release of zymogen granules, which fuse with lysosomes, leading to impaired autophagy. Lysosomal cathepsin B causes premature trypsinogen activation and release of cathepsin B and trypsin into the cytoplasm. The released cathepsin B acts on the RIP3-RIP1-MLKL signaling pathway to promote RIP3-RIP1 necroptosis. It also leads to the release of cytochrome-c from the mitochondria, which activates caspase-3 and cell apoptosis. (3) They trigger ER stress and UPR, leading to CHOP expression and cell apoptosis. Necrotic or necroptotic cellular death liberate DAMPS. For details, see text.

3.1.5. Inflammation

Premature activation of digestive enzymes that occurs early in pancreatitis is not sufficient to explain several aspects of AP. This is because the inflammatory response in pancreatitis was reported to be independent of trypsinogen activation [188]. Therefore, other mechanisms, such as NF-kB and inflammasome activation, are now considered key pathogenic mechanisms in both acute and chronic pancreatitis [30,207,208]. Findings in models of chronic pancreatitis are quite interesting. The severity of fibrosis and the NF-kB activation of chronic inflammation are not mitigated in cathepsin-B-deficient- and trypsinogen-7-deficient mice, suggesting that inflammation is not dependent on trypsin activation in both chronic pancreatitis and AP [209].
The NF-kB implication. Like in most inflammatory conditions, the activation of NF-κB is an early event during pancreatitis observed within minutes after the initiation of the disease due to the constitutive presence of N-FκB in the cytoplasm of acinar cells before the initiation of AP [210,211]. Trypsinogen and NF-κB activation are independent from each other, but they follow similar kinetics [209], possibly due to their common activation by the intracellular Ca++ signaling [212,213]. However, models of experimental pancreatitis suggest an additional, complex involvement of NF-kB beyond the pro-inflammatory role. In fact, data suggest that NF-kB may even protect acinar cells [214,215,216]. Moreover, mice overexpressing active IKKβ kinase showed chronic infiltration of immune cells without acute pancreatitis, but administration of cerulein led to more severe pancreatitis [217]. These findings indicate that constitutive activation of NF-κB leads to an infiltration of immune cells, but pancreatitis only develops after an additional external noxious stimulant.
Another important transcriptional component in acinar cells is AP1 (activator protein 1). It controls pancreatic differentiation, cell death, and inflammation. Mice heterozygous for the orphan nuclear receptor NR5A2 develop an AP1-dependent pre-inflammatory state similar to early acute pancreatitis [218]. Interestingly, NF-kB and AP1 activity vary according to the etiology of pancreatitis. In cerulein models, an activation of both factors was described [219]. On the other hand, ethanol metabolites can either positively or negatively regulate NF-kB and AP1 depending on the presence of oxidative or non-oxidative alcohol metabolites in the pancreas [26,220]. Direct inhibition of NF-kB by certain agents, such as the peroxisome proliferator activator receptor gamma (PPAR-γ) ligand, pyrrolidine dithiocarbamate (PDTC), and calpain I inhibitor, can ameliorate experimental AP, but the clinical significance is still unknown [221].
The role of DAMPs. As a result of cell damage caused by injured acinar cells, damage-related molecular patterns (DAMPs) can be released that may aggravate pancreatic injury, leading to Systemic Inflammatory Response Syndrome (SIRS) [222,223]. This stage of hyper-inflammation is followed by a compensatory anti-inflammatory response syndrome (CARS), which is related to immunosuppression and is characterized by an overproduction of anti-inflammatory cytokines, such as TGF-β, IL-4, and IL-10 [224]. IL-10 inhibits the STAT3 pathway and the production of inflammatory cytokines [225]. The use of insulin-like growth factor 1 and IL-4, which enhance IL-10 production, have attenuated the damage in experimental AP [226,227]. However, it should be stressed that during CARS, patients with acute pancreatitis are susceptible to developing infection of pancreatic necrosis [228].
This idea was recently questioned, and a new approach was proposed where both SIRS and CARS start early and develop in parallel, as shown in severe pancreatitis induced by partial duct ligation with cerulein stimulation. Pancreatic macrophages promote inflammation and simultaneously induce a Th2-cell-mediated response via IL-18. The pro-inflammatory Th1 response was scarcely detectable in concert with the absence of IL-12, a cytokine released by M1-macrophage that regulates Th1 response. Regulatory T-cells were increased and anti-inflammatory M2 macrophages were dominant, while M1-macrophages were identified only in the necrotic areas. Inhibition of the NLRP3 inflammasome reduced both SIRS and CARS. Interestingly, both pathways are regulated by the NLRP3-inflammasome-derived IL-18 [229]. These findings, however, are not in agreement with data from other cerulein models of AP [230] and patients with severe acute pancreatitis [231], where IL-12 is detected and is a predictor of disease severity. This discrepancy may be due to different mechanisms in macrophage stimulation. In vitro activated macrophages after co-incubation with acinar cells did not secrete IL-12, in contrast to macrophage activation with LPS, which does induce IL-12 secretion [232].
DAMPs have a critical role in pancreatic inflammation. High mobility group box 1 (HMGB1) is a nuclear molecule constitutively expressed in almost every cell. HMGB1 may translocate to the cytosol under stress, and then it is released into the extracellular space where it functions as a DAMP with the ability to trigger inflammatory mediators [233]. The circulating HMGB1 levels in AP are increased and correlate with the severity of the disease both in humans and in experimental animal models [234,235,236,237]. In addition, the inhibition of HMGB1 protects from injury in models of AP [238,239,240,241,242]. In addition to HMGGB1, damaged pancreatic acinar cells release different intracellular contents, such as DNA, ATP, and heat shock protein 70 (HSP70), increasing NF-kB activation through TLR4 activation. ATP released by damaged cells also interacts with the purinergic receptor P2x7, inducing mitochondrial dysfunction. This is followed by intracellular K+-depletion, which results in NLRP3 assembly, caspase-1 activation, and IL1β and IL18 secretion [234]. Moreover, the stimulation of intracellular nucleotide-binding oligomerization domain 1 (NOD1) by translocated bacteria from the gut microbiota is a crucial element to aggravate the inflammatory process in the pancreas.
The effects of DAMPs on macrophages were also investigated. Stimulator of interferon genes (STING) activation in macrophages by DNA derived from damaged acinar cells led to the overproduction of pro-inflammatory cytokines by macrophages in experimental AP [243]. In the absence of STING, macrophages did not overproduce cytokines, indicating a direct link between acinar cell DAMPs and the generation of proinflammatory cytokines.
The role of inflammasome activation. Patients with AP have elevated serum levels of pro-inflammatory cytokines, such as IL1β, TNFα, IL6, and IL18 [244]. The precursor forms of IL1β and IL18 cytokines are converted into an active form through the NLRP3 inflammasome. Two signals are required for the activation of inflammasomes. The first signal upregulates the inflammasome mRNA by NF-kB and the second signal initiates the activation of pro-caspase-1. The release of cathepsins from phagosomes into the cytosol may act as the second signal in inflammasome activation [245]. The NLRP3 inflammasome is activated during AP, and its components are required for pancreatic injury. The absence of caspase-1, caspase recruitment domain (ASC), or NLRP3 significantly reduced edema and inflammation in AP [246]. Another study using NLRP3-deficient mice found suppression of IL1β and prevention of the inflammatory cascade [247]. TLR4 involvement in the induction of AP has been reported. Administration of lactate to block TLR4 reduced the activation of NLRP3 inflammasome [248]. This finding agrees with clinical data showing an anti-inflammatory effect of Ringer’s lactate solution used as a fluid replacement in patients with AP [249,250]. Other TLR4 modulators, such as carbon monoxide, produced similar results, indicating a clear role of TLR4 and the NLRP3 inflammasome in AP [251,252]. The NLRP3 inflammasome is also implicated in the development of lung injury secondary to pancreatitis through exosomal release. The plasma-derived exosomes trigger NLRP3 inflammasome activation and pyroptosis in alveolar macrophages, leading to ppulmmonary dysfunction during AP [253]. A human study confirmed animal data. The report confirmed the presence of increased levels of AIM2 and NLRP3 inflammasomes in the early course of AP. Furthermore, AIM2 expression was increased in patients who developed moderate or severe AP [254].
Other factors connecting acinar cell damage and inflammation have been described. Histone deacetylase (HDAC) is one of these factors, as HDAC activity has been demonstrated to play a crucial role in the regulation of inflammation in AP. HDAC inhibition reduced trypsinogen activation, inflammation, and tissue damage in experimental AP [255]. Additionally, inhibition of Sulfiredoxin-1 (Srxn1) expression was reported to increase the production of ROS and induction of apoptosis. Inhibition also promoted inflammation by accumulating M1 macrophages and neutrophils in AP. Overexpression of Srxn1 reduced ROS and apoptosis in acinar cells [256].
Details of inflammatory mechanisms in AP have been published [29].

3.1.6. Role of the Immune System

Early in the course of AP, the pancreas is infiltrated by inflammatory cells. Macrophages and neutrophils are the first to reach the organ and contribute to the pancreatic damage phagocytosing necrotic tissue [257]. Pancreatitis is a sterile inflammation, and pathogen-associated molecular patterns (PAMPs) play no role, at least in the early phases. The activation of immune cells is mediated by damage-associated molecular patterns (DAMPs) that arise from acinar cell necrosis. DAMPs increase the nuclear translocation of the NF-kB family within infiltrating immune cells, leading to enhancement of the cytokine storm [198]. Injured acinar cells also release chemokines that recruit immune cells within minutes after the onset of disease into the site of injury [221,258] The monocyte chemoattractant protein 1 (MCP1) facilitates monocyte trafficking, while macrophage inflammatory protein 2α (MIP2α) and CXC chemokine ligand 1 (CXCL1) recruit macrophages and neutrophils [259,260]. Inhibition of chemokines and their receptors prevents pancreatic and distant organ injury in animal models [261,262,263]. Increased serum MCP1 levels correlate with severe acute pancreatitis in humans [264]. Neutrophilic NADP oxidase promoted oxidative stress and increased intra-acinar trypsinogen activation [265,266]. The infiltration of immune cells has been also associated with the prognosis of pancreatitis [267,268,269]. Macrophage infiltration better correlates with pancreatic damage and necrosis than the number of neutrophils because macrophages are required for the removal of necrosis and thus ameliorate pancreatic damage. Phagocytosing macrophages are found in almost all models of AP and CP [198,267,270,271]. Depletion of macrophages decreases disease severity and protects mice from cerulein-induced pancreatitis [263,267]. Macrophages also produce large amounts of IL1β, which is released by the gasdermin D pore complex from the cytosol into the extracellular space. Consequently, the cell undergoes pyroptotic cell death [272,273,274]. Macrophages at distant organs are also activated and contribute to distant organ damage in AP, although the mechanisms of distal organ injury have not been fully elucidated [264].
The role of neutrophil extracellular traps (NETs). Activated neutrophils use nuclear DNA and histones to form extracellular web-like structures called neutrophil extracellular traps (NETs) that participate in microorganism eradication. However, NETs can cause ductal obstruction, activate pro-inflammatory signals, and prematurely activate trypsinogen [275]. During experimental AP, NETs are produced in the pancreas, regulating organ inflammation and injury. NET levels are also increased in plasma from patients with AP [150,276].
NETs act as a double-edged sword, regulating, on the one hand, the protective innate immune response, but also precipitating in epithelial and tissue injury [277,278]. NETs are also associated with the severity of AP [279]. In septic AP, NETs kill invading pathogens [275] but also activate trypsinogen, mostly through the STAT-3 pathway [280]. NETs may also cause severe damage to other organs, such as the lungs, blood vessels, and kidneys [281,282]. NETs are implicated in thrombosis and participate, therefore, in the hypercoagulability observed in the incipient stage of severe acute pancreatitis [280]. Additionally, NET formation increases macrophage recruitment by releasing chemokines.
As a consequence of these abnormalities in immune mechanisms, a paradoxical period of immunosuppression develops during AP. In patients with mild and moderate AP, there is a reduction in HLA-DR expression within the first few days of the disease, but this returns to near-normal levels within the first week [283]. In contrast, in patients with severe AP, the reduction of HLA-DR expression persists for a long period, and these patients may develop infectious complications [284]. A recent study reported increased activity of the PD1/PD-L1 system in severe pancreatitis, which was more pronounced in patients who develop secondary infectious complications [285]. Another serious consequence of the abnormal immune mechanisms in AP is the impairment of the intestinal barrier, leading to systemic bacterial translocation. In most AP patients, gut barrier failure occurs at the point of hospital admission [286], before the onset of multi-organ failure. Increased intestinal permeability has been observed even in mild disease, although it is more pronounced in severe disease [286,287,288], when patients are more vulnerable to infections [289]. A review of immunopathological abnormalities in AP has very recently been published [289].
Taken together, early protease activation as well as NF-κB and inflammasome activation are essential mechanisms of pancreatitis. These events occur in parallel during disease evolution and strongly influence each other. Recently, it has become clear that not only the activation of proteases and NF-κB play a critical role, but also the type of cell where these events take place. Pancreatitis is no longer a disease of acinar cells alone [26]. The role of inflammation and immunity is summarized in Figure 3.
DAMPs, such as HMGB1, HSP70, and ATP, are crucial for the promotion of acute pancreatitis and, indirectly, of chronic pancreatitis. They activate NF-kB through TLR4 and upregulate the mRNA and protein expression of NLRP3, leading to the assembly of the NLPR3 inflammasome in association with ASC and pro-caspase-1. In addition, bacterial translocation sustains inflammation through the NOD1. NLRP3 and casp-1 are activated by either ATP via its interaction with P2X7 and the resultant intracellular K+-depletion or the ROS produced by NETs. Maturation of pro-IL1β and pro-IL18 leads to IL1β and IL18 secretion and inflammation. Some intracellular DAMPS, such as mitochondrial DNA, initiate NLRP3 inflammasome assembly and activation. Additionally, TLR9 senses intracellular bacteria and mtDNA with subsequent activation of NF-κB. Chemokines produced by injured acinar cells recruit and activate macrophages and neutrophils at the site of injury. Cytokines, such as TGFβ1, produced by macrophages activate PSCs to produce extracellular matrix initiating fibrosis in chronic pancreatitis. Neutrophils also release NETs, aggravating inflammation.

3.1.7. Exosomes and AP

Exosomes are vesicles secreted by various living cells that contain RNA and proteins (30–100 nm in size). In experimental pancreatitis, the number and content of the exosomes released to the peripheral blood from the pancreas are significantly increased. As mentioned before, pancreatic exosomes can reach the lung through circulation, where they are phagocytosed by alveolar macrophages, changing their phenotype from M2 to M1, which in turn aggravates the lung injury caused by AP [290]. Plasma-derived exosomes may activate lung NLRP3 inflammasomes to induce pyroptosis of alveolar macrophages in AP. Inhibition of these exosomes represses the pyroptosis of alveolar macrophages, attenuating the AP-induced lung damage [253]. The interplay between acinar cells and macrophages has been confirmed through analysis of microRNAs found in exosomes. Acinar cells activate macrophages through exosomes released in AP, which in turn promote acinar cell injury via apoptosis, necrosis, and autophagy [291]. Exosomes derived from different cells are not always detrimental in AP. Thus, exosomes derived from bone marrow mesenchymal stem cells (MSCs) have a protective effect on AP [292]. The specificity of exosomes in different cells and tissues should be further investigated [125].

3.1.8. Genetic Mutations

Several gene mutations are implicated in the pathogenesis of acute pancreatitis, such as mutations in protease serine 1, serine protease inhibitor Kazal type 1, chymotrypsin C, the cystic fibrosis transmembrane conductance regulator (CFTR), claudin 2, and calcium-sensing receptor genes [293]. Human genetic data indicate that premature activation or misfolding of pancreatic proteases play a central role in the onset of pancreatitis and progression to chronic pancreatitis [26]. A detailed presentation of these genetic mutations is beyond the scope of this review, but a paper is available that elegantly summarizes the genetics of acute pancreatitis [294].

4. Chronic Pancreatitis

4.1. Pathophysiology of Chronic Pancreatitis (CP)

Multiple mechanisms are involved in the pathogenesis of CP. Repeated insults to the pancreas by alcohol or tobacco or any other factor may lead to recurrent attacks of AP, which in turn activate pancreatic stellate cells (PSCs) and initiate fibrogenesis, ultimately resulting in chronic fibrosing pancreatitis. Interestingly, these recurrent attacks very frequently cause histopathological abnormalities in the pancreas in many patients, who remain asymptomatic, and only a few experience clinical disease [31,295,296,297]. After repeated episodes, areas of pancreatic necrosis are replaced by fibrotic tissue [298,299]. However, pancreatic necrosis is uncommon in patients with classic chronic pancreatitis.
Another theory is the “two-hit” model. After infiltration of the pancreas by macrophages and neutrophils during an episode of AP and the activation of pancreatic stellate cells (PSCs), a second continuous insult, such as alcohol and tobacco and their metabolites, will promote fibrosis through activated immune cells [300,301,302,303]. However, most patients do not proceed to CP despite continuous use of these toxic factors. Earlier reports pointed to the role of ROS production by acinar cells and the subsequent activation of NF-kB [267]. ROS can promote the fusion of lysosomes and zymogen granules and the premature activation of trypsinogen [304,305]. This theory totally ignores the critical role of ROS production by macrophages. Finally, a ductal dysfunction has been connected to CP pathogenesis. Reduction of secretion of bicarbonate-rich fluid [306] favors the formation of protein plugs and ductal obstruction. Protein plugs are indeed described in chronic pancreatitis [298], but it is not clear if they are the cause or the result of CP.
The drawbacks of all of the theories presented above indicate that external factors are not sufficient to explain the development of CP. Whatever the pathogenesis of CP might be, its development leads to pancreatic exocrine insufficiency, diabetes mellitus, and an immune response that results in nerve abnormalities and chronic pain [307,308].
Mutations in several genes, such as the cationic and ionic trypsinogen or the pancreas-specific protease elastase 3b (CELA3B), have been implicated in the pathophysiology of CP to complete some of the missing points [16,26,309,310]. Therefore, it seems that a different “two-hit” model, where the first step is an underlying mutation and the second is the effect of toxic external factors, is a more comprehensive pathogenetic theory. The influence of external risk factors is very strong. In patients with CP, the pooled prevalence of alcohol as a risk factor is 65% compared to the risk factor of 61% of tobacco [17]. Smoking or alcohol abstinence reduces the risk of disease progression [311,312]. Alcohol toxicity is due to its metabolites [313,314] that cause the microcirculatory disturbances, which in turn mediate pancreatic acinar cell injury, resulting in fibrosis and chronic disease [172,315]. Recently, macrophages and pancreatic stellate cells are in the center of extensive investigation to clarify their role in the pathogenesis of CP [316].

4.2. The Role of Macrophages

Macrophages are the main inflammatory cells implicated in CP fibrosis [317]. Fibrogenesis in CP is induced when macrophages and other inflammatory cells are attracted by tissue damage and infiltrate the pancreas [318]. Necrosis and apoptosis of acinar cells can activate macrophages. Macrophages produce transforming growth factor-β1 (TGF-β1), platelet-derived growth factor (PDGF), and connective tissue growth factor (CTGF), which initiate activation and proliferation of the resident PSCs, which are transformed in mmyofibroblast-like cells (PMF) [319,320]. Activated macrophages create a positive feedback cycle through PSCs to secrete more cytokines [198,321]. Activated PSCs in turn induce M2 macrophages that play a significant role in angiogenesis and promote tissue fibrosis [322,323,324,325].

4.3. The Role of PSCs

PSCs are able to oxidize alcohol to acetaldehyde, leading to the generation of ROS and oxidative stress. As mentioned before, alcohol consumption damages the intestinal barrier and increases circulating levels of lipopolysaccharide, oxidized low-density lipoproteins, and TNF-α. All of them are potential activators of PSC in concert with TGF-β1 [326,327]. The resultant myofibroblasts secrete increased amounts of extracellular matrix proteins, thus mediating pancreatic fibrosis [328,329,330]. Mice overexpressing TGF-β1 develop spontaneous pancreatic fibrosis, indicating that TGF-β1 activates de novo PSCs [331]. Activated PSCs can also secrete CTGF, IL-1/16, and endothelin-1 (ET-1) and further promote the activation of PSCs through autocrine and paracrine signaling, which forms a vicious cycle [332]. At present, TGF-β is the strongest activator of PSCs. This effect is achieved by regulating the Smad2/3 signaling pathway [333,334]. This mechanism also influences the phosphorylation of three subtypes of the MAPK family, including c-Jun amino-terminal kinase (JNK), p38, and ERK [335,336,337]. All of these signals work in parallel, leading to pancreatic fibrosis [338,339,340]. While TGF-β1 is critical for promoting matrix deposition by myofibroblasts, it fails to induce PMF proliferation, in contrast to PDGF, wwwwhich drove proliferation of PMFs isolated from CP patients [341]. PDGF also stimulated the production of a matrix with reduced potency compared with TGF-β1 [342]. PDGF itself is not capable of initial activation of PSCs [343].
CTGF is the third activator of PSCs. In CP specimens, CTGF and TGF-β1 were increased more than 20-fold [344]. It should be noted that CTGF expression in the pancreas is controlled by several cytokines, such as TGF-β1, Activin-A, PDGF, and TNF-α [345,346].
Additional external activators of PSCs are TNF-α, IL-1β, and Cyclooxygenase-2 (COX2). Incubation of PMFs with TNF-α increased α-SMA expression [342,347,348]. The COX2 downstream product Prostaglandin E2 stimulated PMF proliferation and the expression of matrix proteins and matrix metalloproteinases [349]. In a different model, ectopic expression of COX-2 in the acinar cells of rodents led to spontaneous CP, with deposition of ECM proteins [350]. PMFs are also sensitive to DAMPs. Rodent PMFs express TLRs 2, 3, 4, and 5, along with co-receptors CD14 and MD2 [351]. TLR2 and 4 are known receptors of high mobility group box 1 (HMGB1), heat shock protein 70 (HSP70), and fibrinogen, indicating that DAMPS are directly implicated in pancreatic fibrosis. In addition to DAMPs, acinar cell damage can directly activate PSCs. After trypsinogen activation, acinar cells liberate a large number of cytokines that activate PSCs and induce fibrosis [352,353,354]. Acinar cells in CP gradually change from columnar to flat and form a ductal structure expressing cytokeratin. This transformation is called acinar ductal metaplasia (ADM), and it is a promoter of fibrosis and an early event of pancreatic adenocarcinoma [355,356].
Two more factors that activate PCSs prove the complex nature of fibrosis development in CP. Hypoxia activates PSCs with the increased release of type I collagen, fibronectin, and vascular endothelial growth factor (VEGF) [357]. As mentioned before, high pressure in the pancreatic duct stimulates Piezo1 channel opening, leading to PCSs’ activation and pressure-induced chronic pancreatitis. This mechanism may explain the fibrosis developed in biliary CP [358]. Details of the activation of PSCs were recently published [359,360].
In addition to the innate response mostly mediated by macrophages, several studies have demonstrated an involvement of adaptive immunity as well [361]. Increased clusters of CD4+ and CD8+ T cells in parallel with increased IL-10 levels have been reported in CP patients compared with healthy individuals [362]. Moreover, chronic pancreatitis specimens had more disease-specific regulatory T-cell subsets [363] and central chemokine receptor 7 (CCR7) positive memory T cells that persisted up to 3 years after pancreatic resection [364].

5. A Brief Synopsis of Forms of Cellular Death in Acute and Chronic Pancreatitis

1.
Apoptosis was the first form of regulated cell death (RCD) to be described [365]. Apoptosis includes both external and internal pathways. The external pathway is initiated by death receptors (such as TNF receptors or Fas receptors) and mediated by the initiator caspase-8. Intrinsic apoptosis is initiated by MOMP, which leads to the release of mitochondrial proteins, such as cytochrome c, and diablo IAP-binding mitochondrial protein (DIABLO, also known as Smac), and subsequent activation of the initiator caspase-9 [366]. Both pathways lead to the activation of executionar caspases and cellular death.
2.
Necroptosis. The regulated process of necrosis is called necroptosis. It is mediated by RIPs and MLKL, as mentioned before. Compared to apoptosis, necroptosis may be a more aggressive mode of cell death. Recent studies indicated that necroptosis may be the main mechanism of acinar cell death in AP [365,367,368].
3.
Pyroptosis is the result of NLRP3 and other inflammasome activation. IL-37 protects against acinar cell pyroptosis in AP [369]. The activation of pyroptosis includes the caspase-1-dependent canonical pathway and the caspase-4/5/11-dependent non-canonical pathway. Caspases-3-7-8, implicated in apoptosis, also participate in the regulation of pyroptosis [370]. Caspases-1-4-5-11 directly cleave the gasdermin D (GSDMD) to produce N-terminal fragments. GSDMD forms pores in the plasma membrane, followed by membrane rupture. It has been proposed that a shift from apoptosis to pyroptosis and necroptosis may explain why some patients with pancreatitis develop the necrotizing form of the disease [229,371].
4.
Ferroptosis is a new RCD pathway that is an iron-dependent form of non-apoptotic cell death first described in 2012. It is induced by accumulation of peroxidized lipids and is regulated by glutathione peroxidase 4 (GPX4) and arachidonic acid lipid oxygenases [372]. Ferroptosis plays an important role in the death of acinar cells, at least in AP, associated with hypertriglyceridemia. NADPH oxidase 2 (NOX2) is a key point in the regulation of ferroptosis. The inhibition of ferroptosis and NOX2 attenuated the inflammatory response in a rodent model of AP and improved the outcome [373].

6. Autophagy in Pancreatitis

Autophagy is the cellular pathway for organelle, lipid, and protein degradation. It is the more efficient recycling machinery in nature [96].
Genetic models targeting autophagy have in part clarified the significant role of this system in the pancreatic pathophysiology. The role of autophagy was investigated in mice with pancreas-specific knockouts of mediators of autophagosome formation, the autophagy-related proteins ATG5 or ATG7. Genetic deletions of ATG5 or ATG7 or of the inhibitor of nuclear factor IκB kinase α (IKKα) result in ER stress and accumulation of dysfunctional mitochondria unable to generate ATP [178,271]. Moreover, the lysosome associated membrane protein 2 (LAMP2) deficiency increased the severity of cerulein pancreatitis [270,374]. Administration of the enhancer of autophagy trehalose significantly reduced trypsinogen activation and necrosis in a murine pancreatitis model [122]. Importantly, tissue from patients with pancreatitis showed abnormalities of autophagy similar to those in murine models [122,374,375,376]. These will be analyzed below. The AP models of IL-22 transgenic mice are a further indication of autophagy involvement in AP, as IL-22 can prevent the formation of autophagosomes through the Beclin-1 pathway, reducing the severity of AP [377]. An important aspect of autophagy in severe AP is the effect of autophagy on the integrity of intestinal barrier. Reduced autophagy in severe AP impairs tight and gap junctions and reduces the function of goblet and Paneth cells, leading to increased bacterial translocation and extra-pancreatic serious manifestations [370,378,379]. The increase in oxidative stress associated with the increased bacterial translocation will aggravate AP-associated lung injury and was attributed to decreased autophagy levels [380]. However, the opposite has also been reported, as excessive autophagy may also be connected to lung injury. It was recently shown that the nuclear translocation of Nrf2 reduced excessive autophagy in severe acute pancreatitis-related acute lung injury via the p62–Kelch-like ECH-associated protein 1 (Keap1)-NF-E2-related factor 2 (Nrf2) signaling pathway in mice [381].
Recent investigations have revealed more associations of autophagy and pancreatitis. Thus, an additional connection between zymogen exocytosis and autophagy has been reported involving SNARE proteins. Syntaxin 2 (STX-2), a SNARE protein of the acinar cell, blocked the fusion of zymogen granules with the plasma membrane and exocytosis and, at the same time, deregulated autophagosome formation by disrupting autophagy-related 16-like 1 protein (Atg16L1), an interaction with the clathrin heavy chain. This interaction is necessary to recruit membranes from acinar plasma membrane for physiologic autophagosome formation [382]. Notably, depletion of another SNARE protein, SNAP23, prevented the induction of AP by reducing trypsin activation of autolysosomes [383].
Xanthohumol (Xn), a natural prenylated chalcone compound isolated from hops, restored autophagy flux by inhibiting the AKT/mTOR pathway in experimental pancreatitis. This was associated with reductions in necrosis, inflammation, oxidative stress, and the severity of pancreatitis [384]. Experiments also indicated that Pancreatic Protein kinase C iota (PKCi) significantly increased pancreatic immune cell infiltration, acinar cell DNA damage, and apoptosis, but reduced sensitivity to cerulein-induced pancreatitis. Prkci deletion in acinar cells resulted in p62 aggregation and loss of autophagic vesicles consistent with the disruption of autophagy [385].
Farnesoid X receptor (FXR) has been also implicated in pancreatitis. FXR is a ligand-activated factor that has an important role in the regulation of glucose, lipid, bile acid, and amino acid metabolism [386]. FXR is also an anti-inflammatory factor in several inflammatory diseases [387,388]. Nuclear FXR was considerably increased in the pancreas of patients with pancreatitis accompanied by a parallel increase in Oxidative Stress Induced Growth Inhibitor 1 (OSGIN1), which is the direct target of FXR in the exocrine pancreas. Deletion of the FXR in acinar cells caused severe pancreatitis, whereas pancreatic overexpression of Osgin1 reduced the severity of pancreatitis. Stimulation of autophagic flux by the FXR-OSGIN1 axis was the mechanism through which FXR-OSGIN1 protected against pancreatitis [389].
A selective autophagic pathway called zymophagy is an early protective mechanism in AP preventing acinar cell death [390,391]. It may be induced by CCK-receptor hyperstimulation and may account for the self-limited form of AP [390].
The protective effect of canonical autophagy was reported in a recent study comparing canonical autophagy with the Ras-related protein Rab9-mediated non-canonical autophagy, which was not protective. These two forms of autophagy antagonize each other. Thus, Rab9 decrease as observed in rodent and human pancreatitis may be a beneficial response to boost canonical autophagy and mitigate disease severity [392].
Autophagy and autolysosomes are additionally involved in trypsinogen activation, as suggested by earlier reports. In a model of AP with atg5 deletion, reduced severity of the disease paralleled reduced trypsinogen activation [393,394], a finding verified by a subsequent study [376]. Trypsinogen activation and the role of autophagy have been reviewed in detail [395,396].

6.1. Autophagy and ER in AP

ER is responsible for the synthesis and folding of proteins, the storage of Ca++, and the regulation of Ca2+ concentration in cells [397]. Endoplasmic reticulum stress (ER stress) develops when the ER is overwhelmed by unfolded and misfolded proteins. Morphological changes in ER indicating ER stress, such as swollen ER, vacuolation, and loss of ribosomes, are observed at the early stage of AP [398,399]. ER is closely associated with autophagy. The major membrane source for the creation of autophagosomes is the rough endoplasmic reticulum, and both the initiation and maturation of autophagosomes have a close relationship with ER [179,400,401]. Autophagy will be interrupted, or the already impaired autophagy will deteriorate after the development of ER stress. Moreover, IL-1β released by macrophages can cause ER stress and liberation of large amounts of Ca2+ from ER into the cytoplasm, leading to both activation of trypsinogen and impaired autophagy in murine pancreatitis [402,403,404]. Alcohol consumption can also induce ER stress that impairs lysosomal proteases and lysosomal membrane proteins, such as LAMP2, leading to deranged autophagy and initiation of AP [405]. The deletion of IκB kinase α (IKKα) gene impaired autophagy and P62 accumulation, leading to ER stress and spontaneous pancreatitis [375]. With P62 gene deletion, all of these damages were mitigated, suggesting that autophagy impairment can indeed cause ER stress [176,406]. The ATG7 gene knockout model showed that autophagosomes are not formed in acinar cells, and autophagy flux is reduced while ER stress is increased [178]. Trehalose, which can increase autophagy activity and restore autophagy flux, reduces ER stress and trypsinogen activation, thus alleviating AP, as mentioned before [122]. In a different murine, it was also shown that reduction of autophagy aggravated AP and increased ER stress [407]. Taken together, these findings indicate that there is a reverse association between autophagy and ER stress. This is further confirmed by evidence suggesting that restoration of ER function could in turn promote autophagy and protect acinar cells. Thus, melatonin administration inhibited the EER stress and promoted autophagy, alleviating AP [408]. Finally, it should be noted that there is a synergy between UPR and autophagic pathways. Both UPR and autophagy aim to restore ER function, as autophagy also degrades misfolded proteins, and the specialized form of reticulophagy removes damaged ER [405,409].

6.2. Autophagy and Mitochondria in AP

Mitochondrial dysfunction can lead to impairment of autophagy through the CypD-related MPTP opening. In some AP animal models, such as the cerulean and bile acid models, mitochondrial dysfunction in acinar cells is moderated through Ca2++-dependent pathways [161]. Ca2++-independent pathways may operate in other models, such as the Arginine-induced model of AP, where the opening of MPTPs is due to the decreased ATP synthase activity [410], while the MPTP opening in alcoholic AP is mediated by the reduction of Nicotinamide adenine dinucleotide (NAD) [354]. Finally, they all lead to continuous opening of the MPTPs, which is controlled by the mitochondria resident protein CypD [122,152]. Inactivation of CypD restores mitochondrial polarity and ATP synthase activity, proving that mitochondria regulate lysosomes and therefore autophagy in the pancreas [122,152,155]. In more detail, in the arginine model, free Arg in the mitochondria of acinar cells increased, and it was degraded through the ornithine pathway. The degradation product reduced ATP synthase, resulting in reduced autophagy, ER stress, and lipid metabolism disorders, ultimately leading to AP [122]. It was recently reported that loss of estrogen-related receptor γ (ERRγ) resulted in mitochondrial dysfunction and further increased autophagosome accumulation and ER stress in acinar cells [411].
In addition, impaired autophagy can also influence mitochondrial function through inefficient mitophagy, the selective autophagy of mitochondria [84,154]. Acinar cell survival depends on the efficient removal of damaged mitochondria. AP in mice induces mitophagy by up-regulating Parkin, an E3 ubiquitin-protein ligase that initiates mitophagy, as mentioned before [179,412]. Normal mitophagy may in part explain the mild course of AP in the majority of patients. However, the deletions of atg5 and atg7 genes inhibit mitophagy and lead to the accumulation of dysfunctional mitochondria [178,271], suggesting that the impaired autophagy observed in AP is finally accompanied by reduced mitophagy [354]. Recently, a new pathway for mitophagy was demonstrated in AP. Alterations of mitochondrial dynamics and subsequent mitochondrial dysfunction were shown early in the acute phase of mild pancreatitis. Moreover, it was shown that the vacuole membrane protein-1 (VMP1) is necessary in mitophagy, as VMP1 downregulation significantly reduced mitochondrial degradation [413]. Overproduction of ROS may also disrupt mitophagy, causing severe AP by activating the AKT/mTOR pathway [414]. So far, the results in human AP agree with the experimental findings.

6.3. Autophagy and Lysosomes in AP

The lysosome contains more than 60 acid hydrolases. It is protected from auto-ddegrradation by a glycocalyx of the membrane [415,416]. The lysosome is considered today to be an important coordinator of signals regulating cell growth, proliferation, and differentiation, in addition to its participation in the final stage of autophagy [417]. Cathepsins are the most important acid hydrolases of the lysosomes [418]. Dysfunction of the lysosomes can block autophagy through three mechanisms. The first mechanism is the impairment of the fusion of lysosomes with autophagosomes due to the defective function of the lysosomal membrane proteins, LAMP-1 and LAMP-2 [270,419,420]. This is an important mechanism of alcohol’s induction of AP and CP. In murine pancreatitis, alcohol reduces LAMP-2 proteins, leading to the accumulation of autophagosomes in acinar cells and a shift from apoptosis to necrosis [374]. Patients with alcoholic pancreatitis also have local LAMP-2 depletion. Abnormal cathepsins are the basis of the second mechanism. Pancreatitis impairs the maturation of cathepsins in lysosomes of acinar cells, resulting in the accumulation of autolysosomes with undegraded material, including zymogen granules [395]. It should be noted that a reduction in enzymatic activities of cathepsins in pancreatic lysosomes has been reported in AP [421]. An imbalance between cathepsin L and cathepsin B may be the underlying reason. As mentioned before, cathepsin B converts trypsinogen to trypsin, while cathepsin L degrades both trypsinogen and trypsin. Inhibition of cathepsin L may therefore lead to increased activity of trypsin and pancreatitis [376]. Inadequate synthesis of lysosomes leading to autophagy reduction and induction of AP is the third mechanism. Transcription factor EB (TFEB) is the central regulator of lysosome synthesis [70] and also a transcriptional factor of several autophagy-related genes [422,423]. TFEB is degraded in the cerulein model of AP, resulting in autophagy impairment [424]. Deletion of tfeb increased the severity of murine AP, while tfeb overexpression attenuated pancreatitis [425]. It should be noted that defective or aging lysosomes are phagocytosed by autophagosomes and fused with normal lysosomes for degradation through a process called lysophagy [426,427].
Recently, the importance of normal cathepsins was demonstrated in a double-knockout (DKO) model of cathepsin deficiencies. Cathepsin B/cathepsin D DKO mice showed cytoplasmic degeneration similar to atg5 KO mice. The autophagy markers LC3 and p62 accumulated, and the numbers of autophagosomes increased in the acinar cells. Moreover, these mice developed CP, indicating the significance not only of cathepsin B but also the significance of the combination with cathepsin D. Single KO mice for either cathepsin were normal [428].

6.4. The Role of miRNAs in Regulating Autophagy of AP

As mentioned above, there is an interplay between different organelles and autophagy in pancreatic acinar cells. Several microRNAs are involved in this interplay. Thus, miR155, miR141, miR-181b, miR-148a, and miR-375 contribute to the inhibition of autophagy initiation by inhibiting the expression of Beclin-1. The repressed expression of ATG12 and p62 and the downregulation of LAMP-2 by miR-148b-3p will also derange autophagy [429].
On the other hand, MiR-92b-3p was reported to attenuate inflammation and autophagy in cells incubated with cerulein by targeting tumor necrosis factor receptor-associated factor-3 (TRAF3) and repressing the p38 pathway [430]. An additional regulation of autophagy by miRNAs is via the calcium/calmodulin-dependent protein kinase II (CAMKII). It mediates the phosphorylation of its substrates in response to cytoplasmic Ca++ increase [431,432]. In addition, CAMKII is auto-phosphorylated after the entry of Ca++ into acinar cells and acquires Ca++-independent activity [433]. CAMKII activity is necessary in one model of AP induced by nicardipine [432]. Pancreatic necrosis parallels the level of CAMKII, which is positively controlled by ATG7, suggesting that there is a connection between impaired autophagy and CAMKII-regulated necrosis in the pathogenesis of AP. The level of miR-30b-5p was negatively correlated with the levels of ATG7, indicating that the well-described impairment of autophagy is associated with low ATG7 and the subsequent necrosis of AP is mediated by the miR-30b-5p/CAMKII pathway [434].

6.5. Interplay of Autophagy and Inflammatory Response in Pancreatitis

Autophagy elimination through deletion of the genes atg5, atg7, lamp-2, or ikkα increases the inflammatory reaction associated with up-regulated production of cytokines, chemokines, and macrophage infiltration of the pancreas, as mentioned before [178,270,271,375]. Both inflammatory (M1) and fibrogenic (M2) macrophages are increased. M1 macrophages predominate in LAMP2-null mice with cerulein pancreatitis, while neutrophils are decreased, indicating a shift towards chronic inflammation [270]. In ATG5-deficient mice, autophagy blockade activates NF-kB, STAT3, and cJun N-terminal kkinases, all of which stimulate production of cytokines by acinar cells [435]. ATG5 deficiency also activates the IκB kinase (IKK)-related kinase (TBK1) and increases the infiltration by neutrophils and T-cells accompanied by PD-L1 upregulation, increased levels of type I interferon (IFN), and the IFN-regulated chemokine CCL5 [436]. As mentioned before, persistent ER stress is observed in mouse models of AP [437] and CP [438]. Autophagy suppression is one of the inducers of ER stress in experimental pancreatitis [178,271,439], but the mechanisms linking ER stress with defective autophagy, and the inflammatory response, are not clarified as of yet [315,406,440,441]. Table 1 summarizes the main associations of autophagy with pancreatitis.

7. Future Perspectives

Most pre-clinical studies focus on the regulation of the increased intra acinar calcium, and many ongoing clinical trials try to identify new agents for the treatment of AP. Despite the fact that pharmacological inhibition of the autophagy process offers a potential therapeutic strategy for AP, ongoing clinical trials are not existent, and only pre-clinical studies offer potential future clinical applications [443,444].
Inhibition of autophagy reduces AP severity [445] and alters the progression of experimental AP in mice [446]. Earlier studies showed that activation of the nuclear factor-κB pathway increases autophagy in pancreatic acinus cells, while inhibition of this pathway ameliorated AP [447].
Interleukin 22, a member of the Interleukin-10 family, is the most widely used agent in animal models. IL-22 is increased in experimental AP and in patients with AP. Administration of IL-22 reduced pancreatic inflammation and improved survival [448]. The protective effect of IL-22 on pancreatitis was mediated via the induction of Bcl-2 and Bcl-XL, which bind to Beclin-1 and subsequently inhibit autophagosome formation and the autophagic pathway [377]. A more recent study indicated that the beneficial effect of IL-22 is due to the activation of the AKT/mTOR pathway and subsequent inhibition of autophagy [449].
Spautin-1, an inhibitor of autophagy, was also shown to ameliorate acute inhibiting impaired autophagy and Ca2+ overload [431]. Moreover, a spautin-1 derivative, spautin-A41, was described as a potent autophagy inhibitor. Mice treated with spautin-A41 were resistant to cerulein-induced pancreatitis due to the inhibition of autophagosome formation [450]. Interestingly, sitagliptin, a dipeptidyl peptidase-4 (DPP4) inhibitor recently associated with autophagy, ameliorated AP-induced acute lung injury. Sitagliptin protection was attributed to the reduction of excessive autophagy through the p62-Keap1-Nrf2 signaling pathway [381].
Clinical trials are required to verify in patients the significance of autophagy modulators.

8. Conclusions

The pathogenesis of both acute and chronic pancreatitis is a complicated process involving several pathways. The traditional theory of premature activation of trypsinogen into the acinar cells has been complemented by various signals in both acinar and ductal cells of the pancreas. Mitochondrial dysfunction and ER stress are prominent features of pancreatitis pathophysiology. Moreover, calcium signaling, exosome abnormalities, and the implication of mechanisms related to inflammation, innate immunity, and genetic predisposition have been clarified. Macrophages are recognized as important mediators of inflammation and innate immunity. MicroRNA regulation of inflammation has also been explored. Fibrosis induction by macrophages and pancreatic stellate cells are prominent characteristics of disease progression towards chronic pancreatitis. Most importantly, the role of autophagy and its specialized forms, such as mitophagy, are now at the center of interest. Autophagy has been associated with both protection and aggravation of experimental and human pancreatitis. It is the common denominator behind practically every mechanism involved in the pathogenesis of pancreatitis and a target for possible therapeutic interventions in this disease.

Author Contributions

Study concept and design (I.T. and E.K.), acquisition of data (E.K.), analysis and interpretation of data (E.K. and A.V.), drafting of the manuscript (I.T. and E.K.), critical revision of the manuscript for important intellectual content (E.K. and I.T.), administrative, technical, or material support (I.T. and A.V.), and study supervision (I.T. and E.K.). All authors have made a significant contribution to this study and have approved the final manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yadav, D.; Lowenfels, A.B. The epidemiology of pancreatitis and pancreatic cancer. Gastroenterology 2013, 144, 1252–1261. [Google Scholar] [CrossRef] [PubMed]
  2. Coté, G.A.; Yadav, D.; Slivka, A.; Hawes, R.H.; Anderson, M.A.; Burton, F.R.; Brand, R.E.; Banks, P.A.; Lewis, M.D.; Disario, J.A.; et al. Alcohol and smoking as risk factors in an epidemiology study of patients with chronic pancreatitis. Clin. Gastroenterol. Hepatol. 2011, 9, 266–273; quiz e27. [Google Scholar] [CrossRef] [PubMed]
  3. Setiawan, V.W.; Monroe, K.; Lugea, A.; Yadav, D.; Pandol, S. Uniting Epidemiology and Experimental Disease Models for Alcohol-Related Pancreatic Disease. Alcohol. Res. 2017, 38, 173–182. [Google Scholar] [PubMed]
  4. Setiawan, V.W.; Pandol, S.J.; Porcel, J.; Wilkens, L.R.; Le Marchand, L.; Pike, M.C.; Monroe, K.R. Prospective Study of Alcohol Drinking, Smoking, and Pancreatitis: The Multiethnic Cohort. Pancreas 2016, 45, 819–825. [Google Scholar] [CrossRef] [PubMed]
  5. Yadav, D.; Hawes, R.H.; Brand, R.E.; Anderson, M.A.; Money, M.E.; Banks, P.A.; Bishop, M.D.; Baillie, J.; Sherman, S.; DiSario, J.; et al. Alcohol consumption, cigarette smoking, and the risk of recurrent acute and chronic pancreatitis. Arch. Intern. Med. 2009, 169, 1035–1045. [Google Scholar] [CrossRef] [PubMed]
  6. Boxhoorn, L.; Voermans, R.P.; Bouwense, S.A.; Bruno, M.J.; Verdonk, R.C.; Boermeester, M.A.; van Santvoort, H.C.; Besselink, M.G. Acute pancreatitis. Lancet 2020, 396, 726–734. [Google Scholar] [CrossRef] [PubMed]
  7. Valdivielso, P.; Ramírez-Bueno, A.; Ewald, N. Current knowledge of hypertriglyceridemic pancreatitis. Eur. J. Intern. Med. 2014, 25, 689–694. [Google Scholar] [CrossRef] [PubMed]
  8. Adiamah, A.; Psaltis, E.; Crook, M.; Lobo, D.N. A systematic review of the epidemiology, pathophysiology and current management of hyperlipidaemic pancreatitis. Clin. Nutr. 2018, 37 Pt A, 1810–1822. [Google Scholar] [CrossRef]
  9. Roberts, S.E.; Morrison-Rees, S.; John, A.; Williams, J.G.; Brown, T.H.; Samuel, D.G. The incidence and aetiology of acute pancreatitis across Europe. Pancreatology 2017, 17, 155–165. [Google Scholar] [CrossRef] [PubMed]
  10. Gonzalez-Perez, A.; Schlienger, R.G.; Rodríguez, L.A. Acute pancreatitis in association with type 2 diabetes and antidiabetic drugs: A population-based cohort study. Diabetes Care 2010, 33, 2580–2585. [Google Scholar] [CrossRef] [PubMed]
  11. Lai, S.W.; Muo, C.H.; Liao, K.F.; Sung, F.C.; Chen, P.C. Risk of acute pancreatitis in type 2 diabetes and risk reduction on anti-diabetic drugs: A population-based cohort study in Taiwan. Am. J. Gastroenterol. 2011, 106, 1697–1704. [Google Scholar] [CrossRef] [PubMed]
  12. Shafqet, M.; Sharzehi, K. Diabetes and the Pancreatobiliary Diseases. Curr. Treat. Options Gastroenterol. 2017, 15, 508–519. [Google Scholar] [CrossRef] [PubMed]
  13. Forsmark, C.E.; Vege, S.S.; Wilcox, C.M. Acute Pancreatitis. N. Engl. J. Med. 2017, 376, 598–599. [Google Scholar] [CrossRef] [PubMed]
  14. Jalaly, N.Y.; Moran, R.A.; Fargahi, F.; Khashab, M.A.; Kamal, A.; Lennon, A.M.; Walsh, C.; Makary, M.A.; Whitcomb, D.C.; Yadav, D.; et al. An Evaluation of Factors Associated With Pathogenic PRSS1, SPINK1, CTFR, and/or CTRC Genetic Variants in Patients With Idiopathic Pancreatitis. Am. J. Gastroenterol. 2017, 112, 1320–1329. [Google Scholar] [CrossRef] [PubMed]
  15. Whitcomb, D.C. Genetic risk factors for pancreatic disorders. Gastroenterology 2013, 144, 1292–1302. [Google Scholar] [CrossRef] [PubMed]
  16. Whitcomb, D.C.; LaRusch, J.; Krasinskas, A.M.; Klei, L.; Smith, J.P.; Brand, R.E.; Neoptolemos, J.P.; Lerch, M.M.; Tector, M.; Sandhu, B.S.; et al. Common genetic variants in the CLDN2 and PRSS1-PRSS2 loci alter risk for alcohol-related and sporadic pancreatitis. Nat. Genet. 2012, 44, 1349–1354. [Google Scholar] [CrossRef] [PubMed]
  17. Sankaran, S.J.; Xiao, A.Y.; Wu, L.M.; Windsor, J.A.; Forsmark, C.E.; Petrov, M.S. Frequency of progression from acute to chronic pancreatitis and risk factors: A meta-analysis. Gastroenterology 2015, 149, 1490–1500.e1. [Google Scholar] [CrossRef] [PubMed]
  18. Whitcomb, D.C.; Frulloni, L.; Garg, P.; Greer, J.B.; Schneider, A.; Yadav, D.; Shimosegawa, T. Chronic pancreatitis: An international draft consensus proposal for a new mechanistic definition. Pancreatology 2016, 16, 218–224. [Google Scholar] [CrossRef] [PubMed]
  19. Kleeff, J.; Whitcomb, D.C.; Shimosegawa, T.; Esposito, I.; Lerch, M.M.; Gress, T.; Mayerle, J.; Drewes, A.M.; Rebours, V.; Akisik, F.; et al. Chronic pancreatitis. Nat. Rev. Dis. Primers. 2017, 3, 17060. [Google Scholar] [CrossRef] [PubMed]
  20. Beyer, G.; Habtezion, A.; Werner, J.; Lerch, M.M.; Mayerle, J. Chronic pancreatitis. Lancet 2020, 396, 499–512. [Google Scholar] [CrossRef] [PubMed]
  21. Vege, S.S.; Chari, S.T. Chronic Pancreatitis. N. Engl. J. Med. 2022, 386, 869–878. [Google Scholar] [CrossRef] [PubMed]
  22. Singh, V.K.; Yadav, D.; Garg, P.K. Diagnosis and Management of Chronic Pancreatitis: A Review. JAMA 2019, 322, 2422–2434. [Google Scholar] [CrossRef] [PubMed]
  23. Jeon, C.Y.; Whitcomb, D.C.; Slivka, A.; Brand, R.E.; Gelrud, A.; Tang, G.; Abberbock, J.; AlKaade, S.; Guda, N.; Mel Wilcox, C.; et al. Lifetime Drinking History of Persons With Chronic Pancreatitis. Alcohol. Alcohol. 2019, 54, 615–624. [Google Scholar] [CrossRef] [PubMed]
  24. Strum, W.B. Abstinence in alcoholic chronic pancreatitis. Effect on pain and outcome. J. Clin. Gastroenterol. 1995, 20, 37–41. [Google Scholar] [CrossRef] [PubMed]
  25. Strum, W.B.; Spiro, H.M. Chronic pancreatitis. Ann. Intern. Med. 1971, 74, 264–277. [Google Scholar] [CrossRef] [PubMed]
  26. Mayerle, J.; Sendler, M.; Hegyi, E.; Beyer, G.; Lerch, M.M.; Sahin-Tóth, M. Genetics, Cell Biology, and Pathophysiology of Pancreatitis. Gastroenterology 2019, 156, 1951–1968.e1. [Google Scholar] [CrossRef] [PubMed]
  27. Masamune, A.; Kotani, H.; Sörgel, F.L.; Chen, J.M.; Hamada, S.; Sakaguchi, R.; Masson, E.; Nakano, E.; Kakuta, Y.; Niihori, T.; et al. Variants That Affect Function of Calcium Channel TRPV6 Are Associated With Early-Onset Chronic Pancreatitis. Gastroenterology 2020, 158, 1626–1641.e8. [Google Scholar] [CrossRef] [PubMed]
  28. Sahin-Tóth, M. Channelopathy of the Pancreas Causes Chronic Pancreatitis. Gastroenterology 2020, 158, 1538–1540. [Google Scholar] [CrossRef] [PubMed]
  29. Habtezion, A.; Gukovskaya, A.S.; Pandol, S.J. Acute Pancreatitis: A Multifaceted Set of Organelle and Cellular Interactions. Gastroenterology 2019, 156, 1941–1950. [Google Scholar] [CrossRef] [PubMed]
  30. Saluja, A.; Dudeja, V.; Dawra, R.; Sah, R.P. Early Intra-Acinar Events in Pathogenesis of Pancreatitis. Gastroenterology 2019, 156, 1979–1993. [Google Scholar] [CrossRef] [PubMed]
  31. Pitchumoni, C.S.; Glasser, M.; Saran, R.M.; Panchacharam, P.; Thelmo, W. Pancreatic fibrosis in chronic alcoholics and nonalcoholics without clinical pancreatitis. Am. J. Gastroenterol. 1984, 79, 382–388. [Google Scholar] [PubMed]
  32. Hori, Y.; Vege, S.S.; Chari, S.T.; Gleeson, F.C.; Levy, M.J.; Pearson, R.K.; Petersen, B.T.; Kendrick, M.L.; Takahashi, N.; Truty, M.J.; et al. Classic chronic pancreatitis is associated with prior acute pancreatitis in only 50% of patients in a large single-institution study. Pancreatology 2019, 19, 224–229. [Google Scholar] [CrossRef] [PubMed]
  33. Ktistakis, N.T. In praise of M. Anselmier who first used the term “autophagie” in 1859. Autophagy 2017, 13, 2015–2017. [Google Scholar] [CrossRef] [PubMed]
  34. Appelmans, F.; Wattiaux, R.; De Duve, C. Tissue fractionation studies. 5. The association of acid phosphatase with a special class of cytoplasmic granules in rat liver. Biochem. J. 1955, 59, 438–445. [Google Scholar] [CrossRef] [PubMed]
  35. de Duve, C. The lysosome turns fifty. Nat. Cell Biol. 2005, 7, 847–849. [Google Scholar] [CrossRef] [PubMed]
  36. Takeshige, K.; Baba, M.; Tsuboi, S.; Noda, T.; Ohsumi, Y. Autophagy in yeast demonstrated with proteinase-deficient mutants and conditions for its induction. J. Cell Biol. 1992, 119, 301–311. [Google Scholar] [CrossRef]
  37. Klionsky, D.J.; Baehrecke, E.H.; Brumell, J.H.; Chu, C.T.; Codogno, P.; Cuervo, A.M.; Debnath, J.; Deretic, V.; Elazar, Z.; Eskelinen, E.L.; et al. A comprehensive glossary of autophagy-related molecules and processes (2nd edition). Autophagy 2011, 7, 1273–1294. [Google Scholar] [CrossRef] [PubMed]
  38. Harnett, M.M.; Pineda, M.A.; Latré de Laté, P.; Eason, R.J.; Besteiro, S.; Harnett, W.; Langsley, G. From Christian de Duve to Yoshinori Ohsumi: More to autophagy than just dining at home. Biomed. J. 2017, 40, 9–22. [Google Scholar] [CrossRef] [PubMed]
  39. Ohsumi, Y. Historical landmarks of autophagy research. Cell Res. 2014, 24, 9–23. [Google Scholar] [CrossRef] [PubMed]
  40. Gonzalez Porras, M.A.; Sieck, G.C.; Mantilla, C.B. Impaired Autophagy in Motor Neurons: A Final Common Mechanism of Injury and Death. Physiology 2018, 33, 211–224. [Google Scholar] [CrossRef] [PubMed]
  41. Corona Velazquez, A.F.; Jackson, W.T. So Many Roads: The Multifaceted Regulation of Autophagy Induction. Mol. Cell Biol. 2018, 38, e00303-18. [Google Scholar] [CrossRef] [PubMed]
  42. Liang, N.; He, Q.; Liu, X.; Sun, H. Multifaceted roles of ATM in autophagy: From nonselective autophagy to selective autophagy. Cell Biochem. Funct. 2019, 37, 177–184. [Google Scholar] [CrossRef] [PubMed]
  43. Zachari, M.; Ganley, I.G. The mammalian ULK1 complex and autophagy initiation. Essays Biochem. 2017, 61, 585–596. [Google Scholar] [CrossRef] [PubMed]
  44. Kim, J.; Kundu, M.; Viollet, B.; Guan, K.L. AMPK and mTOR regulate autophagy through direct phosphorylation of Ulk1. Nat. Cell Biol. 2011, 13, 132–141. [Google Scholar] [CrossRef]
  45. Russell, R.C.; Tian, Y.; Yuan, H.; Park, H.W.; Chang, Y.Y.; Kim, J.; Kim, H.; Neufeld, T.P.; Dillin, A.; Guan, K.L. ULK1 induces autophagy by phosphorylating Beclin-1 and activating VPS34 lipid kinase. Nat. Cell Biol. 2013, 15, 741–750. [Google Scholar] [CrossRef] [PubMed]
  46. Ren, H.; Zhao, F.; Zhang, Q.; Huang, X.; Wang, Z. Autophagy and skin wound healing. Burn. Trauma. 2022, 10, tkac003. [Google Scholar] [CrossRef] [PubMed]
  47. Birgisdottir, Å.B.; Johansen, T. Autophagy and endocytosis—Interconnections and interdependencies. J. Cell Sci. 2020, 133, jcs228114. [Google Scholar] [CrossRef]
  48. Levine, B.; Sinha, S.; Kroemer, G. Bcl-2 family members: Dual regulators of apoptosis and autophagy. Autophagy 2008, 4, 600–606. [Google Scholar] [CrossRef] [PubMed]
  49. Kang, R.; Zeh, H.J.; Lotze, M.T.; Tang, D. The Beclin 1 network regulates autophagy and apoptosis. Cell Death Differ. 2011, 18, 571–580. [Google Scholar] [CrossRef]
  50. Parzych, K.R.; Klionsky, D.J. An overview of autophagy: Morphology, mechanism, and regulation. Antioxid. Redox Signal. 2014, 20, 460–473. [Google Scholar] [CrossRef]
  51. Dikic, I.; Elazar, Z. Mechanism and medical implications of mammalian autophagy. Nat. Rev. Mol. Cell Biol. 2018, 19, 349–364. [Google Scholar] [CrossRef] [PubMed]
  52. Cheng, X.; Ma, X.; Ding, X.; Li, L.; Jiang, X.; Shen, Z.; Chen, S.; Liu, W.; Gong, W.; Sun, Q. Pacer Mediates the Function of Class III PI3K and HOPS Complexes in Autophagosome Maturation by Engaging Stx17. Mol. Cell. 2017, 65, 1029–1043.e5. [Google Scholar] [CrossRef] [PubMed]
  53. Cheng, X.; Ma, X.; Zhu, Q.; Song, D.; Ding, X.; Li, L.; Jiang, X.; Wang, X.; Tian, R.; Su, H.; et al. Pacer Is a Mediator of mTORC1 and GSK3-TIP60 Signaling in Regulation of Autophagosome Maturation and Lipid Metabolism. Mol. Cell 2019, 73, 788–802.e7. [Google Scholar] [CrossRef]
  54. Sheng, J.Q.; Wang, M.R.; Fang, D.; Liu, L.; Huang, W.J.; Tian, D.A.; He, X.X.; Li, P.Y. LncRNA NBR2 inhibits tumorigenesis by regulating autophagy in hepatocellular carcinoma. Biomed. Pharmacother. 2021, 133, 111023. [Google Scholar] [CrossRef] [PubMed]
  55. Di Malta, C.; Cinque, L.; Settembre, C. Transcriptional Regulation of Autophagy: Mechanisms and Diseases. Front. Cell Dev. Biol. 2019, 7, 114. [Google Scholar] [CrossRef] [PubMed]
  56. Galluzzi, L.; Green, D.R. Autophagy-Independent Functions of the Autophagy Machinery. Cell 2019, 177, 1682–1699. [Google Scholar] [CrossRef] [PubMed]
  57. Shahrabi, S.; Paridar, M.; Zeinvand-Lorestani, M.; Jalili, A.; Zibara, K.; Abdollahi, M.; Khosravi, A. Autophagy regulation and its role in normal and malignant hematopoiesis. J. Cell Physiol. 2019, 234, 21746–21757. [Google Scholar] [CrossRef] [PubMed]
  58. Thorburn, A. Autophagy and disease. J. Biol. Chem. 2018, 293, 5425–5430. [Google Scholar] [CrossRef] [PubMed]
  59. Nakamura, S.; Yoshimori, T. New insights into autophagosome-lysosome fusion. J. Cell Sci. 2017, 130, 1209–1216. [Google Scholar] [CrossRef] [PubMed]
  60. Katsuragi, Y.; Ichimura, Y.; Komatsu, M. p62/SQSTM1 functions as a signaling hub and an autophagy adaptor. FEBS J. 2015, 282, 4672–4678. [Google Scholar] [CrossRef]
  61. Itakura, E.; Kishi-Itakura, C.; Mizushima, N. The hairpin-type tail-anchored SNARE syntaxin 17 targets to autophagosomes for fusion with endosomes/lysosomes. Cell 2012, 151, 1256–1269. [Google Scholar] [CrossRef] [PubMed]
  62. Matsui, T.; Jiang, P.; Nakano, S.; Sakamaki, Y.; Yamamoto, H.; Mizushima, N. Autophagosomal YKT6 is required for fusion with lysosomes independently of syntaxin 17. J. Cell Biol. 2018, 217, 2633–2645. [Google Scholar] [CrossRef] [PubMed]
  63. Hegedűs, K.; Takáts, S.; Boda, A.; Jipa, A.; Nagy, P.; Varga, K.; Kovács, A.L.; Juhász, G. The Ccz1-Mon1-Rab7 module and Rab5 control distinct steps of autophagy. Mol. Biol. Cell. 2016, 27, 3132–3142. [Google Scholar] [CrossRef] [PubMed]
  64. Vaites, L.P.; Paulo, J.A.; Huttlin, E.L.; Harper, J.W. Systematic Analysis of Human Cells Lacking ATG8 Proteins Uncovers Roles for GABARAPs and the CCZ1/MON1 Regulator C18orf8/RMC1 in Macroautophagic and Selective Autophagic Flux. Mol. Cell Biol. 2017, 38, e00392-17. [Google Scholar] [CrossRef] [PubMed]
  65. Ferguson, C.J.; Lenk, G.M.; Meisler, M.H. Defective autophagy in neurons and astrocytes from mice deficient in PI(3,5)P2. Hum. Mol. Genet. 2009, 18, 4868–4878. [Google Scholar] [CrossRef] [PubMed]
  66. Gómez-Virgilio, L.; Silva-Lucero, M.D.; Flores-Morelos, D.S.; Gallardo-Nieto, J.; Lopez-Toledo, G.; Abarca-Fernandez, A.M.; Zacapala-Gómez, A.E.; Luna-Muñoz, J.; Montiel-Sosa, F.; Soto-Rojas, L.O.; et al. Autophagy: A Key Regulator of Homeostasis and Disease: An Overview of Molecular Mechanisms and Modulators. Cells 2022, 11, 2262. [Google Scholar] [CrossRef] [PubMed]
  67. Peña-Llopis, S.; Vega-Rubin-de-Celis, S.; Schwartz, J.C.; Wolff, N.C.; Tran, T.A.; Zou, L.; Xie, X.J.; Corey, D.R.; Brugarolas, J. Regulation of TFEB and V-ATPases by mTORC1. EMBO J. 2011, 30, 3242–3258. [Google Scholar] [CrossRef] [PubMed]
  68. Yu, L.; McPhee, C.K.; Zheng, L.; Mardones, G.A.; Rong, Y.; Peng, J.; Mi, N.; Zhao, Y.; Liu, Z.; Wan, F.; et al. Termination of autophagy and reformation of lysosomes regulated by mTOR. Nature 2010, 465, 942–946. [Google Scholar] [CrossRef] [PubMed]
  69. Puertollano, R.; Ferguson, S.M.; Brugarolas, J.; Ballabio, A. The complex relationship between TFEB transcription factor phosphorylation and subcellular localization. EMBO J. 2018, 37, e98804. [Google Scholar] [CrossRef] [PubMed]
  70. Settembre, C.; Di Malta, C.; Polito, V.A.; Garcia Arencibia, M.; Vetrini, F.; Erdin, S.; Erdin, S.U.; Huynh, T.; Medina, D.; Colella, P.; et al. TFEB links autophagy to lysosomal biogenesis. Science 2011, 332, 1429–1433. [Google Scholar] [CrossRef] [PubMed]
  71. Yan, S. Role of TFEB in Autophagy and the Pathogenesis of Liver Diseases. Biomolecules 2022, 12, 672. [Google Scholar] [CrossRef] [PubMed]
  72. Barthez, M.; Poplineau, M.; Elrefaey, M.; Caruso, N.; Graba, Y.; Saurin, A.J. Human ZKSCAN3 and Drosophila M1BP are functionally homologous transcription factors in autophagy regulation. Sci. Rep. 2020, 10, 9653. [Google Scholar] [CrossRef] [PubMed]
  73. Pan, H.Y.; Valapala, M. Role of the Transcriptional Repressor Zinc Finger with KRAB and SCAN Domains 3 (ZKSCAN3) in Retinal Pigment Epithelial Cells. Cells 2021, 10, 2504. [Google Scholar] [CrossRef] [PubMed]
  74. Kim, S.; Eun, H.S.; Jo, E.K. Roles of Autophagy-Related Genes in the Pathogenesis of Inflammatory Bowel Disease. Cells 2019, 8, 77. [Google Scholar] [CrossRef] [PubMed]
  75. Rudnik, S.; Damme, M. The lysosomal membrane-export of metabolites and beyond. FEBS J. 2021, 288, 4168–4182. [Google Scholar] [CrossRef] [PubMed]
  76. Codogno, P.; Mehrpour, M.; Proikas-Cezanne, T. Canonical and non-canonical autophagy: Variations on a common theme of self-eating? Nat. Rev. Mol. Cell Biol. 2011, 13, 7–12. [Google Scholar] [CrossRef] [PubMed]
  77. Heckmann, B.L.; Boada-Romero, E.; Cunha, L.D.; Magne, J.; Green, D.R. LC3-Associated Phagocytosis and Inflammation. J. Mol. Biol. 2017, 429, 3561–3576. [Google Scholar] [CrossRef] [PubMed]
  78. Martinez, J. LAP it up, fuzz ball: A short history of LC3-associated phagocytosis. Curr. Opin. Immunol. 2018, 55, 54–61. [Google Scholar] [CrossRef] [PubMed]
  79. Münz, C. Non-canonical Functions of Macroautophagy Proteins During Endocytosis by Myeloid Antigen Presenting Cells. Front. Immunol. 2018, 9, 2765. [Google Scholar] [CrossRef]
  80. Heckmann, B.L.; Green, D.R. LC3-associated phagocytosis at a glance. J. Cell Sci. 2019, 132, jcs222984. [Google Scholar] [CrossRef] [PubMed]
  81. Durgan, J.; Florey, O. A new flavor of cellular Atg8-family protein lipidation—alternative conjugation to phosphatidylserine during CASM. Autophagy 2021, 17, 2642–2644. [Google Scholar] [CrossRef] [PubMed]
  82. Durgan, J.; Lystad, A.H.; Sloan, K.; Carlsson, S.R.; Wilson, M.I.; Marcassa, E.; Ulferts, R.; Webster, J.; Lopez-Clavijo, A.F.; Wakelam, M.J.; et al. Non-canonical autophagy drives alternative ATG8 conjugation to phosphatidylserine. Mol. Cell. 2021, 81, 2031–2040.e8. [Google Scholar] [CrossRef] [PubMed]
  83. Foerster, E.G.; Mukherjee, T.; Cabral-Fernandes, L.; Rocha, J.D.B.; Girardin, S.E.; Philpott, D.J. How autophagy controls the intestinal epithelial barrier. Autophagy 2022, 18, 86–103. [Google Scholar] [CrossRef] [PubMed]
  84. Youle, R.J.; Narendra, D.P. Mechanisms of mitophagy. Nat. Rev. Mol. Cell Biol. 2011, 12, 9–14. [Google Scholar] [CrossRef] [PubMed]
  85. Frank, M.; Duvezin-Caubet, S.; Koob, S.; Occhipinti, A.; Jagasia, R.; Petcherski, A.; Ruonala, M.O.; Priault, M.; Salin, B.; Reichert, A.S. Mitophagy is triggered by mild oxidative stress in a mitochondrial fission dependent manner. Biochim. Biophys. Acta 2012, 1823, 2297–2310. [Google Scholar] [CrossRef]
  86. Ryter, S.W.; Bhatia, D.; Choi, M.E. Autophagy: A Lysosome-Dependent Process with Implications in Cellular Redox Homeostasis and Human Disease. Antioxid. Redox Signal. 2019, 30, 138–159. [Google Scholar] [CrossRef] [PubMed]
  87. Durcan, T.M.; Fon, E.A. The three ‘P’s of mitophagy: PARKIN, PINK1, and post-translational modifications. Genes Dev. 2015, 29, 989–999. [Google Scholar] [CrossRef] [PubMed]
  88. Narendra, D.P.; Jin, S.M.; Tanaka, A.; Suen, D.F.; Gautier, C.A.; Shen, J.; Cookson, M.R.; Youle, R.J. PINK1 is selectively stabilized on impaired mitochondria to activate Parkin. PLoS Biol. 2010, 8, e1000298. [Google Scholar] [CrossRef] [PubMed]
  89. Kondapalli, C.; Kazlauskaite, A.; Zhang, N.; Woodroof, H.I.; Campbell, D.G.; Gourlay, R.; Burchell, L.; Walden, H.; Macartney, T.J.; Deak, M.; et al. PINK1 is activated by mitochondrial membrane potential depolarization and stimulates Parkin E3 ligase activity by phosphorylating Serine 65. Open Biol. 2012, 2, 120080. [Google Scholar] [CrossRef] [PubMed]
  90. Koyano, F.; Okatsu, K.; Kosako, H.; Tamura, Y.; Go, E.; Kimura, M.; Kimura, Y.; Tsuchiya, H.; Yoshihara, H.; Hirokawa, T.; et al. Ubiquitin is phosphorylated by PINK1 to activate parkin. Nature 2014, 510, 162–166. [Google Scholar] [CrossRef] [PubMed]
  91. Geisler, S.; Holmström, K.M.; Skujat, D.; Fiesel, F.C.; Rothfuss, O.C.; Kahle, P.J.; Springer, W. PINK1/Parkin-mediated mitophagy is dependent on VDAC1 and p62/SQSTM1. Nat. Cell Biol. 2010, 12, 119–131. [Google Scholar] [CrossRef] [PubMed]
  92. Chen, Y.; Dorn, G.W., 2nd. PINK1-phosphorylated mitofusin 2 is a Parkin receptor for culling damaged mitochondria. Science. 2013, 340, 471–475. [Google Scholar] [CrossRef] [PubMed]
  93. Lazarou, M.; Sliter, D.A.; Kane, L.A.; Sarraf, S.A.; Wang, C.; Burman, J.L.; Sideris, D.P.; Fogel, A.I.; Youle, R.J. The ubiquitin kinase PINK1 recruits autophagy receptors to induce mitophagy. Nature 2015, 524, 309–314. [Google Scholar] [CrossRef]
  94. Richter, B.; Sliter, D.A.; Herhaus, L.; Stolz, A.; Wang, C.; Beli, P.; Zaffagnini, G.; Wild, P.; Martens, S.; Wagner, S.A.; et al. Phosphorylation of OPTN by TBK1 enhances its binding to Ub chains and promotes selective autophagy of damaged mitochondria. Proc. Natl. Acad. Sci. USA 2016, 113, 4039–4044. [Google Scholar] [CrossRef] [PubMed]
  95. Moore, A.S.; Holzbaur, E.L. Dynamic recruitment and activation of ALS-associated TBK1 with its target optineurin are required for efficient mitophagy. Proc. Natl. Acad. Sci. USA 2016, 113, E3349–E3358. [Google Scholar] [CrossRef] [PubMed]
  96. Yao, R.Q.; Ren, C.; Xia, Z.F.; Yao, Y.M. Organelle-specific autophagy in inflammatory diseases: A potential therapeutic target underlying the quality control of multiple organelles. Autophagy 2021, 17, 385–401. [Google Scholar] [CrossRef] [PubMed]
  97. Xu, Y.; Shen, J.; Ran, Z. Emerging views of mitophagy in immunity and autoimmune diseases. Autophagy 2020, 16, 3–17. [Google Scholar] [CrossRef] [PubMed]
  98. Ke, P.Y. Mitophagy in the Pathogenesis of Liver Diseases. Cells 2020, 9, 831. [Google Scholar] [CrossRef] [PubMed]
  99. Alim Al-Bari, A.; Ito, Y.; Thomes, P.G.; Menon, M.B.; García-Macia, M.; Fadel, R.; Stadlin, A.; Peake, N.; Faris, M.E.; Eid, N.; et al. Emerging mechanistic insights of selective autophagy in hepatic diseases. Front. Pharmacol. 2023, 14, 1149809. [Google Scholar] [CrossRef] [PubMed]
  100. Onishi, M.; Yamano, K.; Sato, M.; Matsuda, N.; Okamoto, K. Molecular mechanisms and physiological functions of mitophagy. EMBO J. 2021, 40, e104705. [Google Scholar] [CrossRef] [PubMed]
  101. Puleston, D.J.; Simon, A.K. Autophagy in the immune system. Immunology 2014, 141, 1–8. [Google Scholar] [CrossRef]
  102. Liu, K.; Zhao, E.; Ilyas, G.; Lalazar, G.; Lin, Y.; Haseeb, M.; Tanaka, K.E.; Czaja, M.J. Impaired macrophage autophagy increases the immune response in obese mice by promoting proinflammatory macrophage polarization. Autophagy 2015, 11, 271–284. [Google Scholar] [CrossRef] [PubMed]
  103. Gutierrez, M.G.; Master, S.S.; Singh, S.B.; Taylor, G.A.; Colombo, M.I.; Deretic, V. Autophagy is a defense mechanism inhibiting BCG and Mycobacterium tuberculosis survival in infected macrophages. Cell 2004, 119, 753–766. [Google Scholar] [CrossRef]
  104. Oh, J.E.; Lee, H.K. Autophagy as an innate immune modulator. Immune. Netw. 2013, 13, 1–9. [Google Scholar] [CrossRef] [PubMed]
  105. Wei, J.; Long, L.; Yang, K.; Guy, C.; Shrestha, S.; Chen, Z.; Wu, C.; Vogel, P.; Neale, G.; Green, D.R.; et al. Autophagy enforces functional integrity of regulatory T cells by coupling environmental cues and metabolic homeostasis. Nat. Immunol. 2016, 17, 277–285. [Google Scholar] [CrossRef] [PubMed]
  106. Peral de Castro, C.; Jones, S.A.; Ní Cheallaigh, C.; Hearnden, C.A.; Williams, L.; Winter, J.; Lavelle, E.C.; Mills, K.H.; Harris, J. Autophagy regulates IL-23 secretion and innate T cell responses through effects on IL-1 secretion. J. Immunol. 2012, 189, 4144–4153. [Google Scholar] [CrossRef] [PubMed]
  107. Yonekawa, T.; Thorburn, A. Autophagy and cell death. Essays Biochem. 2013, 55, 105–117. [Google Scholar]
  108. Nikoletopoulou, V.; Markaki, M.; Palikaras, K.; Tavernarakis, N. Crosstalk between apoptosis, necrosis and autophagy. Biochim Biophys. Acta 2013, 1833, 3448–3459. [Google Scholar] [CrossRef] [PubMed]
  109. Mariño, G.; Niso-Santano, M.; Baehrecke, E.H.; Kroemer, G. Self-consumption: The interplay of autophagy and apoptosis. Nat. Rev. Mol. Cell Biol. 2014, 15, 81–94. [Google Scholar] [CrossRef] [PubMed]
  110. Ni, H.M.; Bockus, A.; Boggess, N.; Jaeschke, H.; Ding, W.X. Activation of autophagy protects against acetaminophen-induced hepatotoxicity. Hepatology 2012, 55, 222–232. [Google Scholar] [CrossRef] [PubMed]
  111. Liang, Q.; Xiao, Y.; Liu, K.; Zhong, C.; Zeng, M.; Xiao, F. Cr(VI)-Induced Autophagy Protects L-02 Hepatocytes from Apoptosis Through the ROS-AKT-mTOR Pathway. Cell Physiol. Biochem. 2018, 51, 1863–1878. [Google Scholar] [CrossRef] [PubMed]
  112. Zhou, B.; Liu, J.; Kang, R.; Klionsky, D.J.; Kroemer, G.; Tang, D. Ferroptosis is a type of autophagy-dependent cell death. Semin. Cancer Biol. 2020, 66, 89–100. [Google Scholar] [CrossRef] [PubMed]
  113. Liu, J.; Kuang, F.; Kroemer, G.; Klionsky, D.J.; Kang, R.; Tang, D. Autophagy-Dependent Ferroptosis: Machinery and Regulation. Cell Chem. Biol. 2020, 27, 420–435. [Google Scholar] [CrossRef] [PubMed]
  114. Hou, W.; Xie, Y.; Song, X.; Sun, X.; Lotze, M.T.; Zeh, H.J., 3rd; Kang, R.; Tang, D. Autophagy promotes ferroptosis by degradation of ferritin. Autophagy 2016, 12, 1425–1428. [Google Scholar] [CrossRef] [PubMed]
  115. Kang, R.; Tang, D. Autophagy and Ferroptosis—What’s the Connection? Curr. Pathobiol. Rep. 2017, 5, 153–159. [Google Scholar] [CrossRef] [PubMed]
  116. Qu, X.; Gao, H.; Tao, L.; Zhang, Y.; Zhai, J.; Song, Y.; Zhang, S. Autophagy inhibition-enhanced assembly of the NLRP3 inflammasome is associated with cisplatin-induced acute injury to the liver and kidneys in rats. J. Biochem. Mol. Toxicol. 2018, 33, e22208. [Google Scholar] [CrossRef] [PubMed]
  117. Chang, Y.P.; Ka, S.M.; Hsu, W.H.; Chen, A.; Chao, L.K.; Lin, C.C.; Hsieh, C.C.; Chen, M.C.; Chiu, H.W.; Ho, C.L.; et al. Resveratrol inhibits NLRP3 inflammasome activation by preserving mitochondrial integrity and augmenting autophagy. J. Cell Physiol. 2015, 230, 1567–1579. [Google Scholar] [CrossRef] [PubMed]
  118. de Lavera, I.; Pavon, A.D.; Paz, M.V.; Oropesa-Avila, M.; de la Mata, M.; Alcocer-Gomez, E.; Garrido-Maraver, J.; Cotan, D.; Alvarez-Cordoba, M.; Sanchez-Alcazar, J.A. The Connections Among Autophagy, Inflammasome and Mitochondria. Curr. Drug Targets. 2017, 18, 1030–1038. [Google Scholar] [CrossRef] [PubMed]
  119. Broz, P.; Dixit, V.M. Inflammasomes: Mechanism of assembly, regulation and signalling. Nat. Rev. Immunol. 2016, 16, 407–420. [Google Scholar] [CrossRef] [PubMed]
  120. Wang, X.; Jiang, L.; Shi, L.; Yao, K.; Sun, X.; Yang, G.; Jiang, L.; Zhang, C.; Wang, N.; Zhang, H.; et al. Zearalenone induces NLRP3-dependent pyroptosis via activation of NF-κB modulated by autophagy in INS-1 cells. Toxicology 2019, 428, 152304. [Google Scholar] [CrossRef] [PubMed]
  121. Wang, Z.; Zhang, S.; Xiao, Y.; Zhang, W.; Wu, S.; Qin, T.; Yue, Y.; Qian, W.; Li, L. NLRP3 Inflammasome and Inflammatory Diseases. Oxid Med. Cell Longev. 2020, 2020, 4063562. [Google Scholar] [CrossRef] [PubMed]
  122. Biczo, G.; Vegh, E.T.; Shalbueva, N.; Mareninova, O.A.; Elperin, J.; Lotshaw, E.; Gretler, S.; Lugea, A.; Malla, S.R.; Dawson, D.; et al. Mitochondrial Dysfunction, Through Impaired Autophagy, Leads to Endoplasmic Reticulum Stress, Deregulated Lipid Metabolism, and Pancreatitis in Animal Models. Gastroenterology 2018, 154, 689–703. [Google Scholar] [CrossRef] [PubMed]
  123. Li, H.; Wu, D.; Zhang, H.; Li, P. New insights into regulatory cell death and acute pancreatitis. Heliyon 2023, 9, e18036. [Google Scholar] [CrossRef] [PubMed]
  124. Zheng, Z.; Ding, Y.X.; Qu, Y.X.; Cao, F.; Li, F. A narrative review of acute pancreatitis and its diagnosis, pathogenetic mechanism, and management. Ann. Transl. Med. 2021, 9, 69. [Google Scholar] [CrossRef] [PubMed]
  125. Guo, X.Y.; Xiao, F.; Li, J.; Zhou, Y.N.; Zhang, W.J.; Sun, B.; Wang, G. Exosomes and pancreatic diseases: Status, challenges, and hopes. Int. J. Biol. Sci. 2019, 15, 1846–1860. [Google Scholar] [CrossRef] [PubMed]
  126. Hegyi, P.; Petersen, O.H. The exocrine pancreas: The acinar-ductal tango in physiology and pathophysiology. Rev. Physiol. Biochem. Pharmacol. 2013, 165, 1–30. [Google Scholar] [CrossRef] [PubMed]
  127. Hegyi, P.; Rakonczay, Z. Insufficiency of electrolyte and fluid secretion by pancreatic ductal cells leads to increased patient risk for pancreatitis. Am. J. Gastroenterol. 2010, 105, 2119–2120. [Google Scholar] [CrossRef] [PubMed]
  128. Czakó, L.; Yamamoto, M.; Otsuki, M. Pancreatic fluid hypersecretion in rats after acute pancreatitis. Dig. Dis. Sci. 1997, 42, 265–272. [Google Scholar] [CrossRef] [PubMed]
  129. Manso, M.A.; San Román, J.I.; de Dios, I.; García, L.J.; López, M.A. Cerulein-induced acute pancreatitis in the rat. Study of pancreatic secretion and plasma VIP and secretin levels. Dig. Dis. Sci. 1992, 37, 364–368. [Google Scholar] [CrossRef] [PubMed]
  130. Renner, I.G.; Wisner, J.R., Jr. Ceruletide-induced acute pancreatitis in the dog and its amelioration by exogenous secretin. Int. J. Pancreatol. 1986, 1, 39–49. [Google Scholar] [CrossRef] [PubMed]
  131. Niederau, C.; Liddle, R.A.; Ferrell, L.D.; Grendell, J.H. Beneficial effects of cholecystokinin-receptor blockade and inhibition of proteolytic enzyme activity in experimental acute hemorrhagic pancreatitis in mice. Evidence for cholecystokinin as a major factor in the development of acute pancreatitis. J. Clin. Investig. 1986, 78, 1056–1063. [Google Scholar] [CrossRef]
  132. Lankisch, P.G.; Göke, B.; Fölsch, U.R.; Winckler, K.; Otto, J.; Creutzfeldt, W. Influence of secretin on the course of acute experimental pancreatitis in rats. Digestion 1983, 26, 187–191. [Google Scholar] [CrossRef]
  133. Lerch, M.M.; Saluja, A.K.; Rünzi, M.; Dawra, R.; Steer, M.L. Luminal endocytosis and intracellular targeting by acinar cells during early biliary pancreatitis in the opossum. J. Clin. Investig. 1995, 95, 2222–2231. [Google Scholar] [CrossRef] [PubMed]
  134. Noble, M.D.; Romac, J.; Vigna, S.R.; Liddle, R.A. A pH-sensitive, neurogenic pathway mediates disease severity in a model of post-ERCP pancreatitis. Gut 2008, 57, 1566–1571. [Google Scholar] [CrossRef] [PubMed]
  135. Romac, J.M.; Shahid, R.A.; Swain, S.M.; Vigna, S.R.; Liddle, R.A. Piezo1 is a mechanically activated ion channel and mediates pressure induced pancreatitis. Nat. Commun. 2018, 9, 1715. [Google Scholar] [CrossRef] [PubMed]
  136. Venglovecz, V.; Rakonczay, Z., Jr.; Ozsvári, B.; Takács, T.; Lonovics, J.; Varró, A.; Gray, M.A.; Argent, B.E.; Hegyi, P. Effects of bile acids on pancreatic ductal bicarbonate secretion in guinea pig. Gut 2008, 57, 1102–1112. [Google Scholar] [CrossRef] [PubMed]
  137. Vigna, S.R.; Shahid, R.A.; Nathan, J.D.; McVey, D.C.; Liddle, R.A. Leukotriene B4 mediates inflammation via TRPV1 in duct obstruction-induced pancreatitis in rats. Pancreas 2011, 40, 708–714. [Google Scholar] [CrossRef] [PubMed]
  138. Wen, L.; Javed, T.A.; Yimlamai, D.; Mukherjee, A.; Xiao, X.; Husain, S.Z. Transient High Pressure in Pancreatic Ducts Promotes Inflammation and Alters Tight Junctions via Calcineurin Signaling in Mice. Gastroenterology 2018, 155, 1250–1263.e5. [Google Scholar] [CrossRef] [PubMed]
  139. Orabi, A.I.; Wen, L.; Javed, T.A.; Le, T.; Guo, P.; Sanker, S.; Ricks, D.; Boggs, K.; Eisses, J.F.; Castro, C.; et al. Targeted inhibition of pancreatic acinar cell calcineurin is a novel strategy to prevent post-ERCP pancreatitis. Cell Mol. Gastroenterol. Hepatol. 2017, 3, 119–128. [Google Scholar] [CrossRef] [PubMed]
  140. Maléth, J.; Venglovecz, V.; Rázga, Z.; Tiszlavicz, L.; Rakonczay, Z., Jr.; Hegyi, P. Non-conjugated chenodeoxycholate induces severe mitochondrial damage and inhibits bicarbonate transport in pancreatic duct cells. Gut 2011, 60, 136–138. [Google Scholar] [CrossRef]
  141. Perides, G.; Laukkarinen, J.M.; Vassileva, G.; Steer, M.L. Biliary acute pancreatitis in mice is mediated by the G-protein-coupled cell surface bile acid receptor Gpbar1. Gastroenterology 2010, 138, 715–725. [Google Scholar] [CrossRef] [PubMed]
  142. Criddle, D.N.; McLaughlin, E.; Murphy, J.A.; Petersen, O.H.; Sutton, R. The pancreas misled: Signals to pancreatitis. Pancreatology 2007, 7, 436–446. [Google Scholar] [CrossRef] [PubMed]
  143. Li, Z.; Xu, C.; Tao, Y.; Liang, Y.; Liang, Q.; Li, J.; Li, R.; Ye, H. Anisodamine alleviates lipopolysaccharide-induced pancreatic acinar cell injury through NLRP3 inflammasome and NF-κB signaling pathway. J. Recept. Signal Transduct. Res. 2020, 40, 58–66. [Google Scholar] [CrossRef] [PubMed]
  144. Pu, W.L.; Bai, R.Y.; Zhou, K.; Peng, Y.F.; Zhang, M.Y.; Hottiger, M.O.; Li, W.H.; Gao, X.M.; Sun, L.K. Baicalein attenuates pancreatic inflammatory injury through regulating MAPK, STAT 3 and NF-κB activation. Int. Immunopharmacol. 2019, 72, 204–210. [Google Scholar] [CrossRef] [PubMed]
  145. Maléth, J.; Hegyi, P. Ca2+ toxicity and mitochondrial damage in acute pancreatitis: Translational overview. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2016, 371, 20150425. [Google Scholar] [CrossRef] [PubMed]
  146. Feng, S.; Wei, Q.; Hu, Q.; Huang, X.; Zhou, X.; Luo, G.; Deng, M.; Lü, M. Research Progress on the Relationship Between Acute Pancreatitis and Calcium Overload in Acinar Cells. Dig. Dis. Sci. 2019, 64, 25–38. [Google Scholar] [CrossRef] [PubMed]
  147. Carreras-Sureda, A.; Pihán, P.; Hetz, C. Calcium signaling at the endoplasmic reticulum: Fine-tuning stress responses. Cell Calcium. 2018, 70, 24–31. [Google Scholar] [CrossRef] [PubMed]
  148. Gerasimenko, J.V.; Gryshchenko, O.; Ferdek, P.E.; Stapleton, E.; Hébert, T.O.; Bychkova, S.; Peng, S.; Begg, M.; Gerasimenko, O.V.; Petersen, O.H. Ca2+ release-activated Ca2+ channel blockade as a potential tool in antipancreatitis therapy. Proc. Natl. Acad. Sci. USA 2013, 110, 13186–13191. [Google Scholar] [CrossRef]
  149. Zhang, S.L.; Yu, Y.; Roos, J.; Kozak, J.A.; Deerinck, T.J.; Ellisman, M.H.; Stauderman, K.A.; Cahalan, M.D. STIM1 is a Ca2+ sensor that activates CRAC channels and migrates from the Ca2+ store to the plasma membrane. Nature 2005, 437, 902–905. [Google Scholar] [CrossRef]
  150. Gukovskaya, A.S.; Pandol, S.J.; Gukovsky, I. New insights into the pathways initiating and driving pancreatitis. Curr. Opin. Gastroenterol. 2016, 32, 429–435. [Google Scholar] [CrossRef]
  151. Lur, G.; Sherwood, M.W.; Ebisui, E.; Haynes, L.; Feske, S.; Sutton, R.; Burgoyne, R.D.; Mikoshiba, K.; Petersen, O.H.; Tepikin, A.V. InsP3 receptors and Orai channels in pancreatic acinar cells: Co-localization and its consequences. Biochem. J. 2011, 436, 231–239. [Google Scholar] [CrossRef] [PubMed]
  152. Mukherjee, R.; Mareninova, O.A.; Odinokova, I.V.; Huang, W.; Murphy, J.; Chvanov, M.; Javed, M.A.; Wen, L.; Booth, D.M.; Cane, M.C.; et al. Mechanism of mitochondrial permeability transition pore induction and damage in the pancreas: Inhibition prevents acute pancreatitis by protecting production of ATP. Gut 2016, 65, 1333–1346. [Google Scholar] [CrossRef] [PubMed]
  153. Elmunzer, B.J.; Serrano, J.; Chak, A.; Edmundowicz, S.A.; Papachristou, G.I.; Scheiman, J.M.; Singh, V.K.; Varadarajulu, S.; Vargo, J.J.; Willingham, F.F.; et al. Rectal indomethacin alone versus indomethacin and prophylactic pancreatic stent placement for preventing pancreatitis after ERCP: Study protocol for a randomized controlled trial. Trials. 2016, 17, 120. [Google Scholar] [CrossRef] [PubMed]
  154. Shalbueva, N.; Mareninova, O.A.; Gerloff, A.; Yuan, J.; Waldron, R.T.; Pandol, S.J.; Gukovskaya, A.S. Effects of oxidative alcohol metabolism on the mitochondrial permeability transition pore and necrosis in a mouse model of alcoholic pancreatitis. Gastroenterology 2013, 144, 437–446.e6. [Google Scholar] [CrossRef] [PubMed]
  155. Gukovsky, I.; Pandol, S.J.; Gukovskaya, A.S. Organellar dysfunction in the pathogenesis of pancreatitis. Antioxid. Redox Signal. 2011, 15, 2699–2710. [Google Scholar] [CrossRef] [PubMed]
  156. Odinokova, I.V.; Sung, K.F.; Mareninova, O.A.; Hermann, K.; Evtodienko, Y.; Andreyev, A.; Gukovsky, I.; Gukovskaya, A.S. Mechanisms regulating cytochrome c release in pancreatic mitochondria. Gut 2009, 58, 431–442. [Google Scholar] [CrossRef] [PubMed]
  157. Kroemer, G.; Galluzzi, L.; Brenner, C. Mitochondrial membrane permeabilization in cell death. Physiol. Rev. 2007, 87, 99–163. [Google Scholar] [CrossRef] [PubMed]
  158. Baines, C.P.; Gutiérrez-Aguilar, M. The still uncertain identity of the channel-forming unit(s) of the mitochondrial permeability transition pore. Cell Calcium. 2018, 73, 121–130. [Google Scholar] [CrossRef] [PubMed]
  159. Bernardi, P.; Rasola, A.; Forte, M.; Lippe, G. The Mitochondrial Permeability Transition Pore: Channel Formation by F-ATP Synthase, Integration in Signal Transduction, and Role in Pathophysiology. Physiol. Rev. 2015, 95, 1111–1155. [Google Scholar] [CrossRef] [PubMed]
  160. Lee, H.; Yoon, Y. Mitochondrial fission and fusion. Biochem. Soc. Trans. 2016, 44, 1725–1735. [Google Scholar] [CrossRef] [PubMed]
  161. Gukovskaya, A.S.; Gukovsky, I. Which way to die: The regulation of acinar cell death in pancreatitis by mitochondria, calcium, and reactive oxygen species. Gastroenterology 2011, 140, 1876–1880. [Google Scholar] [CrossRef] [PubMed]
  162. Haanes, K.A.; Novak, I. ATP storage and uptake by isolated pancreatic zymogen granules. Biochem. J. 2010, 429, 303–311. [Google Scholar] [CrossRef] [PubMed]
  163. Burgoyne, R.D.; Morgan, A. Secretory granule exocytosis. Physiol. Rev. 2003, 83, 581–632. [Google Scholar] [CrossRef] [PubMed]
  164. Voronina, S.G.; Barrow, S.L.; Simpson, A.W.; Gerasimenko, O.V.; da Silva Xavier, G.; Rutter, G.A.; Petersen, O.H.; Tepikin, A.V. Dynamic changes in cytosolic and mitochondrial ATP levels in pancreatic acinar cells. Gastroenterology 2010, 138, 1976–1987. [Google Scholar] [CrossRef] [PubMed]
  165. Chvanov, M.; De Faveri, F.; Moore, D.; Sherwood, M.W.; Awais, M.; Voronina, S.; Sutton, R.; Criddle, D.N.; Haynes, L.; Tepikin, A.V. Intracellular rupture, exocytosis and actin interaction of endocytic vacuoles in pancreatic acinar cells: Initiating events in acute pancreatitis. J. Physiol. 2018, 596, 2547–2564. [Google Scholar] [CrossRef] [PubMed]
  166. de Brito, O.M.; Scorrano, L. An intimate liaison: Spatial organization of the endoplasmic reticulum-mitochondria relationship. EMBO J. 2010, 29, 2715–2723. [Google Scholar] [CrossRef] [PubMed]
  167. Hayashi, T.; Rizzuto, R.; Hajnoczky, G.; Su, T.P. MAM: More than just a housekeeper. Trends Cell Biol. 2009, 19, 81–88. [Google Scholar] [CrossRef] [PubMed]
  168. Cárdenas, C.; Miller, R.A.; Smith, I.; Bui, T.; Molgó, J.; Müller, M.; Vais, H.; Cheung, K.H.; Yang, J.; Parker, I.; et al. Essential regulation of cell bioenergetics by constitutive InsP3 receptor Ca2+ transfer to mitochondria. Cell 2010, 142, 270–283. [Google Scholar] [CrossRef] [PubMed]
  169. De Smedt, H.; Verkhratsky, A.; Muallem, S. Ca(2+) signaling mechanisms of cell survival and cell death: An introduction. Cell Calcium. 2011, 50, 207–210. [Google Scholar] [CrossRef] [PubMed]
  170. Ron, D.; Walter, P. Signal integration in the endoplasmic reticulum unfolded protein response. Nat. Rev. Mol. Cell Biol. 2007, 8, 519–529. [Google Scholar] [CrossRef] [PubMed]
  171. Wu, J.S.; Li, W.M.; Chen, Y.N.; Zhao, Q.; Chen, Q.F. Endoplasmic reticulum stress is activated in acute pancreatitis. J. Dig. Dis. 2016, 17, 295–303. [Google Scholar] [CrossRef] [PubMed]
  172. Lugea, A.; Gerloff, A.; Su, H.Y.; Xu, Z.; Go, A.; Hu, C.; French, S.W.; Wilson, J.S.; Apte, M.V.; Waldron, R.T.; et al. The Combination of Alcohol and Cigarette Smoke Induces Endoplasmic Reticulum Stress and Cell Death in Pancreatic Acinar Cells. Gastroenterology 2017, 153, 1674–1686. [Google Scholar] [CrossRef]
  173. Hetz, C.; Chevet, E.; Oakes, S.A. Proteostasis control by the unfolded protein response. Nat. Cell Biol. 2015, 17, 829–838. [Google Scholar] [CrossRef] [PubMed]
  174. Zhao, Q.; Tang, X.; Huang, J.; Li, J.; Chen, Q.; Sun, Y.; Wu, J. Melatonin Attenuates Endoplasmic Reticulum Stress in Acute Pancreatitis. Pancreas 2018, 47, 884–891. [Google Scholar] [CrossRef] [PubMed]
  175. Lugea, A.; Tischler, D.; Nguyen, J.; Gong, J.; Gukovsky, I.; French, S.W.; Gorelick, F.S.; Pandol, S.J. Adaptive unfolded protein response attenuates alcohol-induced pancreatic damage. Gastroenterology 2011, 140, 987–997. [Google Scholar] [CrossRef]
  176. Barrera, K.; Stanek, A.; Okochi, K.; Niewiadomska, Z.; Mueller, C.; Ou, P.; John, D.; Alfonso, A.E.; Tenner, S.; Huan, C. Acinar cell injury induced by inadequate unfolded protein response in acute pancreatitis. World J. Gastrointest. Pathophysiol. 2018, 9, 37–46. [Google Scholar] [CrossRef] [PubMed]
  177. Kim, I.; Xu, W.; Reed, J.C. Cell death and endoplasmic reticulum stress: Disease relevance and therapeutic opportunities. Nat. Rev. Drug Discov. 2008, 7, 1013–1030. [Google Scholar] [CrossRef] [PubMed]
  178. Antonucci, L.; Fagman, J.B.; Kim, J.Y.; Todoric, J.; Gukovsky, I.; Mackey, M.; Ellisman, M.H.; Karin, M. Basal autophagy maintains pancreatic acinar cell homeostasis and protein synthesis and prevents ER stress. Proc. Natl. Acad. Sci. USA 2015, 112, E6166–E6174. [Google Scholar] [CrossRef] [PubMed]
  179. Gukovskaya, A.S.; Gukovsky, I.; Algül, H.; Habtezion, A. Autophagy, Inflammation, and Immune Dysfunction in the Pathogenesis of Pancreatitis. Gastroenterology 2017, 153, 1212–1226. [Google Scholar] [CrossRef] [PubMed]
  180. Richardson, C.E.; Kooistra, T.; Kim, D.H. An essential role for XBP-1 in host protection against immune activation in C. elegans. Nature 2010, 463, 1092–1095. [Google Scholar] [CrossRef] [PubMed]
  181. Aoi, K.; Nishio, A.; Okazaki, T.; Takeo, M.; Masuda, M.; Fukui, T.; Uchida, K.; Okazaki, K. Inhibition of the dephosphorylation of eukaryotic initiation factor 2α ameliorates murine experimental pancreatitis. Pancreatology 2019, 19, 548–556. [Google Scholar] [CrossRef] [PubMed]
  182. Kapuy, O.; Márton, M.; Bánhegyi, G.; Vinod, P.K. Multiple system-level feedback loops control life-and-death decisions in endoplasmic reticulum stress. FEBS Lett. 2020, 594, 1112–1123. [Google Scholar] [CrossRef] [PubMed]
  183. Xu, C.; Bailly-Maitre, B.; Reed, J.C. Endoplasmic reticulum stress: Cell life and death decisions. J. Clin. Investig. 2005, 115, 2656–2664. [Google Scholar] [CrossRef] [PubMed]
  184. Jin, H.Z.; Yang, X.J.; Zhao, K.L.; Mei, F.C.; Zhou, Y.; You, Y.D.; Wang, W.X. Apocynin alleviates lung injury by suppressing NLRP3 inflammasome activation and NF-κB signaling in acute pancreatitis. Int. Immunopharmacol. 2019, 75, 105821. [Google Scholar] [CrossRef] [PubMed]
  185. Chen, J.C.; Wu, M.L.; Huang, K.C.; Lin, W.W. HMG-CoA reductase inhibitors activate the unfolded protein response and induce cytoprotective GRP78 expression. Cardiovasc. Res. 2008, 80, 138–150. [Google Scholar] [CrossRef] [PubMed]
  186. Lee, P.J.; Modha, K.; Chua, T.; Chak, A.; Jang, D.; Lopez, R.; Gougol, A.; Papachristou, G.I.; Stevens, T. Association of Statins With Decreased Acute Pancreatitis Severity: A Propensity Score Analysis. J. Clin. Gastroenterol. 2018, 52, 742–746. [Google Scholar] [CrossRef] [PubMed]
  187. Chiari, H. Über die Selbstverdauung des menschlichen Pankreas. Zeitschrift für Heilkunde 1896, 17, 69–96. [Google Scholar]
  188. Dawra, R.; Sah, R.P.; Dudeja, V.; Rishi, L.; Talukdar, R.; Garg, P.; Saluja, A.K. Intra-acinar trypsinogen activation mediates early stages of pancreatic injury but not inflammation in mice with acute pancreatitis. Gastroenterology 2011, 141, 2210–2217.e2. [Google Scholar] [CrossRef] [PubMed]
  189. Talukdar, R.; Sareen, A.; Zhu, H.; Yuan, Z.; Dixit, A.; Cheema, H.; George, J.; Barlass, U.; Sah, R.; Garg, S.K.; et al. Release of Cathepsin B in Cytosol Causes Cell Death in Acute Pancreatitis. Gastroenterology 2016, 151, 747–758.e5. [Google Scholar] [CrossRef] [PubMed]
  190. Liu, Y.; Liu, T.; Lei, T.; Zhang, D.; Du, S.; Girani, L.; Qi, D.; Lin, C.; Tong, R.; Wang, Y. RIP1/RIP3-regulated necroptosis as a target for multifaceted disease therapy (Review). Int. J. Mol. Med. 2019, 44, 771–786. [Google Scholar] [CrossRef] [PubMed]
  191. Louhimo, J.; Steer, M.L.; Perides, G. Necroptosis Is an Important Severity Determinant and Potential Therapeutic Target in Experimental Severe Pancreatitis. Cell Mol. Gastroenterol. Hepatol. 2016, 2, 519–535. [Google Scholar] [CrossRef] [PubMed]
  192. Han, J.; Zhong, C.Q.; Zhang, D.W. Programmed necrosis: Backup to and competitor with apoptosis in the immune system. Nat. Immunol. 2011, 12, 1143–1149. [Google Scholar] [CrossRef] [PubMed]
  193. He, S.; Wang, L.; Miao, L.; Wang, T.; Du, F.; Zhao, L.; Wang, X. Receptor interacting protein kinase-3 determines cellular necrotic response to TNF-alpha. Cell. 2009, 137, 1100–1111. [Google Scholar] [CrossRef] [PubMed]
  194. Wang, G.; Qu, F.Z.; Li, L.; Lv, J.C.; Sun, B. Necroptosis: A potential, promising target and switch in acute pancreatitis. Apoptosis. 2016, 21, 121–129. [Google Scholar] [CrossRef] [PubMed]
  195. Harris, P.A.; Berger, S.B.; Jeong, J.U.; Nagilla, R.; Bandyopadhyay, D.; Campobasso, N.; Capriotti, C.A.; Cox, J.A.; Dare, L.; Dong, X.; et al. Discovery of a First-in-Class Receptor Interacting Protein 1 (RIP1) Kinase Specific Clinical Candidate (GSK2982772) for the Treatment of Inflammatory Diseases. J. Med. Chem. 2017, 60, 1247–1261. [Google Scholar] [CrossRef] [PubMed]
  196. Ren, Y.; Su, Y.; Sun, L.; He, S.; Meng, L.; Liao, D.; Liu, X.; Ma, Y.; Liu, C.; Li, S.; et al. Discovery of a Highly Potent, Selective, and Metabolically Stable Inhibitor of Receptor-Interacting Protein 1 (RIP1) for the Treatment of Systemic Inflammatory Response Syndrome. J. Med. Chem. 2017, 60, 972–986. [Google Scholar] [CrossRef] [PubMed]
  197. Zhan, X.; Wan, J.; Zhang, G.; Song, L.; Gui, F.; Zhang, Y.; Li, Y.; Guo, J.; Dawra, R.K.; Saluja, A.K.; et al. Elevated intracellular trypsin exacerbates acute pancreatitis and chronic pancreatitis in mice. Am. J. Physiol. Gastrointest. Liver Physiol. 2019, 316, G816–G825. [Google Scholar] [CrossRef] [PubMed]
  198. Sendler, M.; Weiss, F.U.; Golchert, J.; Homuth, G.; van den Brandt, C.; Mahajan, U.M.; Partecke, L.I.; Döring, P.; Gukovsky, I.; Gukovskaya, A.S.; et al. Cathepsin B-Mediated Activation of Trypsinogen in Endocytosing Macrophages Increases Severity of Pancreatitis in Mice. Gastroenterology 2018, 154, 704–718.e10. [Google Scholar] [CrossRef] [PubMed]
  199. Gea-Sorlí, S.; Closa, D. Role of macrophages in the progression of acute pancreatitis. World J. Gastrointest. Pharmacol. Ther. 2010, 1, 107–111. [Google Scholar] [CrossRef]
  200. Aghdassi, A.A.; John, D.S.; Sendler, M.; Weiss, F.U.; Reinheckel, T.; Mayerle, J.; Lerch, M.M. Cathepsin D regulates cathepsin B activation and disease severity predominantly in inflammatory cells during experimental pancreatitis. J. Biol. Chem. 2018, 293, 1018–1029. [Google Scholar] [CrossRef] [PubMed]
  201. Whitcomb, D.C.; Gorry, M.C.; Preston, R.A.; Furey, W.; Sossenheimer, M.J.; Ulrich, C.D.; Martin, S.P.; Gates, L.K., Jr.; Amann, S.T.; Toskes, P.P.; et al. Hereditary pancreatitis is caused by a mutation in the cationic trypsinogen gene. Nat. Genet. 1996, 14, 141–145. [Google Scholar] [CrossRef] [PubMed]
  202. Nguyen, T.D.; Moody, M.W.; Steinhoff, M.; Okolo, C.; Koh, D.S.; Bunnett, N.W. Trypsin activates pancreatic duct epithelial cell ion channels through proteinase-activated receptor-2. J. Clin. Investig. 1999, 103, 261–269. [Google Scholar] [CrossRef] [PubMed]
  203. Alvarez, C.; Regan, J.P.; Merianos, D.; Bass, B.L. Protease-activated receptor-2 regulates bicarbonate secretion by pancreatic duct cells in vitro. Surgery 2004, 136, 669–676. [Google Scholar] [CrossRef] [PubMed]
  204. Laukkarinen, J.M.; Weiss, E.R.; van Acker, G.J.; Steer, M.L.; Perides, G. Protease-activated receptor-2 exerts contrasting model-specific effects on acute experimental pancreatitis. J. Biol. Chem. 2008, 283, 20703–20712. [Google Scholar] [CrossRef] [PubMed]
  205. Namkung, W.; Han, W.; Luo, X.; Muallem, S.; Cho, K.H.; Kim, K.H.; Lee, M.G. Protease-activated receptor 2 exerts local protection and mediates some systemic complications in acute pancreatitis. Gastroenterology 2004, 126, 1844–1859. [Google Scholar] [CrossRef] [PubMed]
  206. Sharma, A.; Tao, X.; Gopal, A.; Ligon, B.; Andrade-Gordon, P.; Steer, M.L.; Perides, G. Protection against acute pancreatitis by activation of protease-activated receptor-2. Am. J. Physiol. Gastrointest. Liver Physiol. 2005, 288, G388–G395. [Google Scholar] [CrossRef] [PubMed]
  207. Gaiser, S.; Daniluk, J.; Liu, Y.; Tsou, L.; Chu, J.; Lee, W.; Longnecker, D.S.; Logsdon, C.D.; Ji, B. Intracellular activation of trypsinogen in transgenic mice induces acute but not chronic pancreatitis. Gut 2011, 60, 1379–1388. [Google Scholar] [CrossRef] [PubMed]
  208. Nikam, A.; Dawra, R.; Saluja, A.; Dudeja, V. Pancreatitis: A Tale of Two Proteases. Gastroenterology 2018, 154, 482–484. [Google Scholar] [CrossRef] [PubMed]
  209. Sah, R.P.; Dudeja, V.; Dawra, R.K.; Saluja, A.K. Cerulein-induced chronic pancreatitis does not require intra-acinar activation of trypsinogen in mice. Gastroenterology 2013, 144, 1076–1085.e2. [Google Scholar] [CrossRef] [PubMed]
  210. Gukovsky, I.; Gukovskaya, A.S.; Blinman, T.A.; Zaninovic, V.; Pandol, S.J. Early NF-kappaB activation is associated with hormone-induced pancreatitis. Am. J. Physiol. 1998, 275, G1402–G1414. [Google Scholar] [CrossRef] [PubMed]
  211. Steinle, A.U.; Weidenbach, H.; Wagner, M.; Adler, G.; Schmid, R.M. NF-kappaB/Rel activation in cerulein pancreatitis. Gastroenterology 1999, 116, 420–430. [Google Scholar] [CrossRef] [PubMed]
  212. Krüger, B.; Albrecht, E.; Lerch, M.M. The role of intracellular calcium signaling in premature protease activation and the onset of pancreatitis. Am. J. Pathol. 2000, 157, 43–50. [Google Scholar] [CrossRef] [PubMed]
  213. Han, B.; Logsdon, C.D. CCK stimulates mob-1 expression and NF-kappaB activation via protein kinase C and intracellular Ca(2+). Am. J. Physiol. Cell Physiol. 2000, 278, C344–C351. [Google Scholar] [CrossRef] [PubMed]
  214. Neuhöfer, P.; Liang, S.; Einwächter, H.; Schwerdtfeger, C.; Wartmann, T.; Treiber, M.; Zhang, H.; Schulz, H.U.; Dlubatz, K.; Lesina, M.; et al. Deletion of IκBα activates RelA to reduce acute pancreatitis in mice through up-regulation of Spi2A. Gastroenterology 2013, 144, 192–201. [Google Scholar] [CrossRef] [PubMed]
  215. Rakonczay, Z., Jr.; Hegyi, P.; Takács, T.; McCarroll, J.; Saluja, A.K. The role of NF-kappaB activation in the pathogenesis of acute pancreatitis. Gut 2008, 57, 259–267. [Google Scholar] [CrossRef] [PubMed]
  216. Koike, Y.; Kanai, T.; Saeki, K.; Nakamura, Y.; Nakano, M.; Mikami, Y.; Yamagishi, Y.; Nakamoto, N.; Ebinuma, H.; Hibi, T. MyD88-dependent interleukin-10 production from regulatory CD11b⁺Gr-1(high) cells suppresses development of acute cerulein pancreatitis in mice. Immunol. Lett. 2012, 148, 172–177. [Google Scholar] [CrossRef] [PubMed]
  217. Aleksic, T.; Baumann, B.; Wagner, M.; Adler, G.; Wirth, T.; Weber, C.K. Cellular immune reaction in the pancreas is induced by constitutively active IkappaB kinase-2. Gut 2007, 56, 227–236. [Google Scholar] [CrossRef] [PubMed]
  218. Cobo, I.; Martinelli, P.; Flández, M.; Bakiri, L.; Zhang, M.; Carrillo-de-Santa-Pau, E.; Jia, J.; Sánchez-Arévalo Lobo, V.J.; Megías, D.; Felipe, I.; et al. Transcriptional regulation by NR5A2 links differentiation and inflammation in the pancreas. Nature 2018, 554, 533–537. [Google Scholar] [CrossRef] [PubMed]
  219. Guo, L.; Sans, M.D.; Hou, Y.; Ernst, S.A.; Williams, J.A. c-Jun/AP-1 is required for CCK-induced pancreatic acinar cell dedifferentiation and DNA synthesis in vitro. Am. J. Physiol. Gastrointest. Liver Physiol. 2012, 302, G1381–G1396. [Google Scholar] [CrossRef] [PubMed]
  220. Gukovskaya, A.S.; Mouria, M.; Gukovsky, I.; Reyes, C.N.; Kasho, V.N.; Faller, L.D.; Pandol, S.J. Ethanol metabolism and transcription factor activation in pancreatic acinar cells in rats. Gastroenterology 2002, 122, 106–118. [Google Scholar] [CrossRef]
  221. Jakkampudi, A.; Jangala, R.; Reddy, B.R.; Mitnala, S.; Nageshwar Reddy, D.; Talukdar, R. NF-κB in acute pancreatitis: Mechanisms and therapeutic potential. Pancreatology 2016, 16, 477–488. [Google Scholar] [CrossRef] [PubMed]
  222. Galluzzi, L.; Vitale, I.; Aaronson, S.A.; Abrams, J.M.; Adam, D.; Agostinis, P.; Alnemri, E.S.; Altucci, L.; Amelio, I.; Andrews, D.W.; et al. Molecular mechanisms of cell death: Recommendations of the Nomenclature Committee on Cell Death 2018. Cell Death Differ. 2018, 25, 486–541. [Google Scholar] [CrossRef] [PubMed]
  223. Gaman, L.; Dragos, D.; Vlad, A.; Robu, G.C.; Radoi, M.P.; Stroica, L.; Badea, M.; Gilca, M. Phytoceuticals in Acute Pancreatitis: Targeting the Balance between Apoptosis and Necrosis. Evid. Based Complement Alternat. Med. 2018, 2018, 5264592. [Google Scholar] [CrossRef] [PubMed]
  224. Lee, B.; Zhao, Q.; Habtezion, A. Immunology of pancreatitis and environmental factors. Curr. Opin. Gastroenterol. 2017, 33, 383–389. [Google Scholar] [CrossRef] [PubMed]
  225. Hutchins, A.P.; Diez, D.; Miranda-Saavedra, D. The IL-10/STAT3-mediated anti-inflammatory response: Recent developments and future challenges. Brief Funct. Genomics. 2013, 12, 489–498. [Google Scholar] [CrossRef] [PubMed]
  226. Lin, R.; Chen, F.; Wen, S.; Teng, T.; Pan, Y.; Huang, H. Interleukin-10 attenuates impairment of the blood-brain barrier in a severe acute pancreatitis rat model. J. Inflamm. 2018, 15, 4. [Google Scholar] [CrossRef] [PubMed]
  227. Warzecha, Z.; Dembinski, A.; Ceranowicz, P.; Konturek, S.J.; Tomaszewska, R.; Stachura, J.; Konturek, P.C. IGF-1 stimulates production of interleukin-10 and inhibits development of caerulein-induced pancreatitis. J. Physiol. Pharmacol. 2003, 54, 575–590. [Google Scholar] [PubMed]
  228. Sharma, D.; Jakkampudi, A.; Reddy, R.; Reddy, P.B.; Patil, A.; Murthy, H.V.V.; Rao, G.V.; Reddy, D.N.; Talukdar, R. Association of Systemic Inflammatory and Anti-inflammatory Responses with Adverse Outcomes in Acute Pancreatitis: Preliminary Results of an Ongoing Study. Dig. Dis. Sci. 2017, 62, 3468–3478. [Google Scholar] [CrossRef] [PubMed]
  229. Sendler, M.; van den Brandt, C.; Glaubitz, J.; Wilden, A.; Golchert, J.; Weiss, F.U.; Homuth, G.; De Freitas Chama, L.L.; Mishra, N.; Mahajan, U.M.; et al. NLRP3 Inflammasome Regulates Development of Systemic Inflammatory Response and Compensatory Anti-Inflammatory Response Syndromes in Mice With Acute Pancreatitis. Gastroenterology 2020, 158, 253–269.e14. [Google Scholar] [CrossRef] [PubMed]
  230. Ahmad, A.; Haas De Mello, A.; Szczesny, B.; Törö, G.; Marcatti, M.; Druzhyna, N.; Liaudet, L.; Tarantini, S.; Salomao, R.; Garcia Soriano, F.; et al. Effects of the Poly(ADP-Ribose) Polymerase Inhibitor Olaparib in Cerulein-Induced Pancreatitis. Shock 2020, 53, 653–665. [Google Scholar] [CrossRef]
  231. Gregorić, P.; Doklestić, K.; Stanković, S.; Sijacki, A.; Karamarković, A.; Radenković, D.; Ivancević, N.; Bajec, D. Interleukin-12 as a predictor of outcome in patients with severe acute pancreatitis. Hepatogastroenterology 2014, 61, 208–211. [Google Scholar] [PubMed]
  232. Iyer, S.; Bawa, E.P.; Tarique, M.; Dudeja, V. Know Thy Enemy-Understanding the Role of Inflammation in Severe Acute Pancreatitis. Gastroenterology 2020, 158, 46–48. [Google Scholar] [CrossRef] [PubMed]
  233. Oppenheim, J.J.; Yang, D. Alarmins: Chemotactic activators of immune responses. Curr. Opin. Immunol. 2005, 17, 359–365. [Google Scholar] [CrossRef] [PubMed]
  234. Ferrero-Andrés, A.; Panisello-Roselló, A.; Roselló-Catafau, J.; Folch-Puy, E. NLRP3 Inflammasome-Mediated Inflammation in Acute Pancreatitis. Int. J. Mol. Sci. 2020, 21, 5386. [Google Scholar] [CrossRef] [PubMed]
  235. Kocsis, A.K.; Szabolcs, A.; Hofner, P.; Takács, T.; Farkas, G.; Boda, K.; Mándi, Y. Plasma concentrations of high-mobility group box protein 1, soluble receptor for advanced glycation end-products and circulating DNA in patients with acute pancreatitis. Pancreatology 2009, 9, 383–391. [Google Scholar] [CrossRef] [PubMed]
  236. Lindström, O.; Tukiainen, E.; Kylänpää, L.; Mentula, P.; Rouhiainen, A.; Puolakkainen, P.; Rauvala, H.; Repo, H. Circulating levels of a soluble form of receptor for advanced glycation end products and high-mobility group box chromosomal protein 1 in patients with acute pancreatitis. Pancreas 2009, 38, e215–e220. [Google Scholar] [CrossRef] [PubMed]
  237. Yasuda, T.; Ueda, T.; Shinzeki, M.; Sawa, H.; Nakajima, T.; Takeyama, Y.; Kuroda, Y. Increase of high-mobility group box chromosomal protein 1 in blood and injured organs in experimental severe acute pancreatitis. Pancreas 2007, 34, 487–488. [Google Scholar] [CrossRef] [PubMed]
  238. Sawa, H.; Ueda, T.; Takeyama, Y.; Yasuda, T.; Shinzeki, M.; Nakajima, T.; Kuroda, Y. Blockade of high mobility group box-1 protein attenuates experimental severe acute pancreatitis. World J. Gastroenterol. 2006, 12, 7666–7670. [Google Scholar] [CrossRef] [PubMed]
  239. Luan, Z.G.; Zhang, X.J.; Yin, X.H.; Ma, X.C.; Zhang, H.; Zhang, C.; Guo, R.X. Downregulation of HMGB1 protects against the development of acute lung injury after severe acute pancreatitis. Immunobiology 2013, 218, 1261–1270. [Google Scholar] [CrossRef] [PubMed]
  240. Yuan, H.; Jin, X.; Sun, J.; Li, F.; Feng, Q.; Zhang, C.; Cao, Y.; Wang, Y. Protective effect of HMGB1 a box on organ injury of acute pancreatitis in mice. Pancreas 2009, 38, 143–148. [Google Scholar] [CrossRef] [PubMed]
  241. Zhang, Z.W.; Zhang, Q.Y.; Zhou, M.T.; Liu, N.X.; Chen, T.K.; Zhu, Y.F.; Wu, L. Antioxidant inhibits HMGB1 expression and reduces pancreas injury in rats with severe acute pancreatitis. Dig. Dis. Sci. 2010, 55, 2529–2536. [Google Scholar] [CrossRef] [PubMed]
  242. Yang, Z.Y.; Ling, Y.; Yin, T.; Tao, J.; Xiong, J.X.; Wu, H.S.; Wang, C.Y. Delayed ethyl pyruvate therapy attenuates experimental severe acute pancreatitis via reduced serum high mobility group box 1 levels in rats. World J. Gastroenterol. 2008, 14, 4546–4550. [Google Scholar] [CrossRef]
  243. Zhao, Q.; Wei, Y.; Pandol, S.J.; Li, L.; Habtezion, A. STING Signaling Promotes Inflammation in Experimental Acute Pancreatitis. Gastroenterology 2018, 154, 1822–1835.e2. [Google Scholar] [CrossRef] [PubMed]
  244. Malmstrøm, M.L.; Hansen, M.B.; Andersen, A.M.; Ersbøll, A.K.; Nielsen, O.H.; Jørgensen, L.N.; Novovic, S. Cytokines and organ failure in acute pancreatitis: Inflammatory response in acute pancreatitis. Pancreas 2012, 41, 271–277. [Google Scholar] [CrossRef] [PubMed]
  245. Guo, H.; Callaway, J.B.; Ting, J.P. Inflammasomes: Mechanism of action, role in disease, and therapeutics. Nat. Med. 2015, 21, 677–687. [Google Scholar] [CrossRef] [PubMed]
  246. Hoque, R.; Sohail, M.; Malik, A.; Sarwar, S.; Luo, Y.; Shah, A.; Barrat, F.; Flavell, R.; Gorelick, F.; Husain, S.; et al. TLR9 and the NLRP3 inflammasome link acinar cell death with inflammation in acute pancreatitis. Gastroenterology 2011, 141, 358–369. [Google Scholar] [CrossRef] [PubMed]
  247. Fu, Q.; Zhai, Z.; Wang, Y.; Xu, L.; Jia, P.; Xia, P.; Liu, C.; Zhang, X.; Qin, T.; Zhang, H. NLRP3 Deficiency Alleviates Severe Acute Pancreatitis and Pancreatitis-Associated Lung Injury in a Mouse Model. Biomed. Res. Int. 2018, 2018, 1294951. [Google Scholar] [CrossRef] [PubMed]
  248. Hoque, R.; Farooq, A.; Ghani, A.; Gorelick, F.; Mehal, W.Z. Lactate reduces liver and pancreatic injury in Toll-like receptor- and inflammasome-mediated inflammation via GPR81-mediated suppression of innate immunity. Gastroenterology 2014, 146, 1763–1774. [Google Scholar] [CrossRef] [PubMed]
  249. Wu, B.U.; Hwang, J.Q.; Gardner, T.H.; Repas, K.; Delee, R.; Yu, S.; Smith, B.; Banks, P.A.; Conwell, D.L. Lactated Ringer’s solution reduces systemic inflammation compared with saline in patients with acute pancreatitis. Clin. Gastroenterol. Hepatol. 2011, 9, 710–717.e1. [Google Scholar] [CrossRef] [PubMed]
  250. de-Madaria, E.; Herrera-Marante, I.; González-Camacho, V.; Bonjoch, L.; Quesada-Vázquez, N.; Almenta-Saavedra, I.; Miralles-Maciá, C.; Acevedo-Piedra, N.G.; Roger-Ibáñez, M.; Sánchez-Marin, C.; et al. Fluid resuscitation with lactated Ringer’s solution vs normal saline in acute pancreatitis: A triple-blind, randomized, controlled trial. United Eur. Gastroenterol. J. 2018, 6, 63–72. [Google Scholar] [CrossRef] [PubMed]
  251. Shen, A.; Kim, H.J.; Oh, G.S.; Lee, S.B.; Lee, S.H.; Pandit, A.; Khadka, D.; Choe, S.K.; Kwak, S.C.; Yang, S.H.; et al. NAD+ augmentation ameliorates acute pancreatitis through regulation of inflammasome signalling. Sci. Rep. 2017, 7, 3006. [Google Scholar] [CrossRef] [PubMed]
  252. Xue, J.; Habtezion, A. Carbon monoxide-based therapy ameliorates acute pancreatitis via TLR4 inhibition. J. Clin. Investig. 2014, 124, 437–447. [Google Scholar] [CrossRef] [PubMed]
  253. Wu, X.B.; Sun, H.Y.; Luo, Z.L.; Cheng, L.; Duan, X.M.; Ren, J.D. Plasma-derived exosomes contribute to pancreatitis-associated lung injury by triggering NLRP3-dependent pyroptosis in alveolar macrophages. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165685. [Google Scholar] [CrossRef] [PubMed]
  254. Algaba-Chueca, F.; de-Madaria, E.; Lozano-Ruiz, B.; Martínez-Cardona, C.; Quesada-Vázquez, N.; Bachiller, V.; Tarín, F.; Such, J.; Francés, R.; Zapater, P.; et al. The expression and activation of the AIM2 inflammasome correlates with inflammation and disease severity in patients with acute pancreatitis. Pancreatology 2017, 17, 364–371. [Google Scholar] [CrossRef] [PubMed]
  255. Hartman, H.; Wetterholm, E.; Thorlacius, H.; Regnér, S. Histone deacetylase regulates trypsin activation, inflammation, and tissue damage in acute pancreatitis in mice. Dig. Dis. Sci. 2015, 60, 1284–1289. [Google Scholar] [CrossRef] [PubMed]
  256. He, J.; Ma, M.; Li, D.; Wang, K.; Wang, Q.; Li, Q.; He, H.; Zhou, Y.; Li, Q.; Hou, X.; et al. Sulfiredoxin-1 attenuates injury and inflammation in acute pancreatitis through the ROS/ER stress/Cathepsin B axis. Cell Death Dis. 2021, 12, 626. [Google Scholar] [CrossRef] [PubMed]
  257. Hoque, R.; Mehal, W.Z. Inflammasomes in pancreatic physiology and disease. Am. J. Physiol. Gastrointest. Liver Physiol. 2015, 308, G643–G651. [Google Scholar] [CrossRef] [PubMed]
  258. Watanabe, T.; Kudo, M.; Strober, W. Immunopathogenesis of pancreatitis. Mucosal. Immunol. 2017, 10, 283–298. [Google Scholar] [CrossRef] [PubMed]
  259. Lugea, A.; Waldron, R.T.; Mareninova, O.A.; Shalbueva, N.; Deng, N.; Su, H.Y.; Thomas, D.D.; Jones, E.K.; Messenger, S.W.; Yang, J.; et al. Human Pancreatic Acinar Cells: Proteomic Characterization, Physiologic Responses, and Organellar Disorders in ex Vivo Pancreatitis. Am. J. Pathol. 2017, 187, 2726–2743. [Google Scholar] [CrossRef] [PubMed]
  260. Griffith, J.W.; Sokol, C.L.; Luster, A.D. Chemokines and chemokine receptors: Positioning cells for host defense and immunity. Annu. Rev. Immunol. 2014, 32, 659–702. [Google Scholar] [CrossRef] [PubMed]
  261. Zhou, G.X.; Zhu, X.J.; Ding, X.L.; Zhang, H.; Chen, J.P.; Qiang, H.; Zhang, H.F.; Wei, Q. Protective effects of MCP-1 inhibitor on a rat model of severe acute pancreatitis. Hepatobiliary Pancreat. Dis. Int. 2010, 9, 201–207. [Google Scholar] [PubMed]
  262. Malla, S.R.; Kärrman Mårdh, C.; Günther, A.; Mahajan, U.M.; Sendler, M.; D’Haese, J.; Weiss, F.U.; Lerch, M.M.; Hansen, M.B.; Mayerle, J. Effect of oral administration of AZD8309, a CXCR2 antagonist, on the severity of experimental pancreatitis. Pancreatology 2016, 16, 761–769. [Google Scholar] [CrossRef] [PubMed]
  263. Saeki, K.; Kanai, T.; Nakano, M.; Nakamura, Y.; Miyata, N.; Sujino, T.; Yamagishi, Y.; Ebinuma, H.; Takaishi, H.; Ono, Y.; et al. CCL2-induced migration and SOCS3-mediated activation of macrophages are involved in cerulein-induced pancreatitis in mice. Gastroenterology 2012, 142, 1010–1020.e9. [Google Scholar] [CrossRef] [PubMed]
  264. Papachristou, G.I. Prediction of severe acute pancreatitis: Current knowledge and novel insights. World J. Gastroenterol. 2008, 14, 6273–6275. [Google Scholar] [CrossRef] [PubMed]
  265. Jakkampudi, A.; Jangala, R.; Reddy, R.; Mitnala, S.; Rao, G.V.; Pradeep, R.; Reddy, D.N.; Talukdar, R. Acinar injury and early cytokine response in human acute biliary pancreatitis. Sci. Rep. 2017, 7, 15276. [Google Scholar] [CrossRef] [PubMed]
  266. Ushio-Fukai, M. Compartmentalization of redox signaling through NADPH oxidase-derived ROS. Antioxid. Redox Signal. 2009, 11, 1289–1299. [Google Scholar] [CrossRef] [PubMed]
  267. Sendler, M.; Dummer, A.; Weiss, F.U.; Krüger, B.; Wartmann, T.; Scharffetter-Kochanek, K.; van Rooijen, N.; Malla, S.R.; Aghdassi, A.; Halangk, W.; et al. Tumour necrosis factor α secretion induces protease activation and acinar cell necrosis in acute experimental pancreatitis in mice. Gut 2013, 62, 430–439. [Google Scholar] [CrossRef] [PubMed]
  268. Gukovsky, I.; Li, N.; Todoric, J.; Gukovskaya, A.; Karin, M. Inflammation, autophagy, and obesity: Common features in the pathogenesis of pancreatitis and pancreatic cancer. Gastroenterology 2013, 144, 1199–1209.e4. [Google Scholar] [CrossRef]
  269. Habtezion, A. Inflammation in acute and chronic pancreatitis. Curr. Opin. Gastroenterol. 2015, 31, 395–399. [Google Scholar] [CrossRef] [PubMed]
  270. Mareninova, O.A.; Sendler, M.; Malla, S.R.; Yakubov, I.; French, S.W.; Tokhtaeva, E.; Vagin, O.; Oorschot, V.; Lüllmann-Rauch, R.; Blanz, J.; et al. Lysosome associated membrane proteins maintain pancreatic acinar cell homeostasis: LAMP-2 deficient mice develop pancreatitis. Cell Mol. Gastroenterol. Hepatol. 2015, 1, 678–694. [Google Scholar] [CrossRef]
  271. Diakopoulos, K.N.; Lesina, M.; Wörmann, S.; Song, L.; Aichler, M.; Schild, L.; Artati, A.; Römisch-Margl, W.; Wartmann, T.; Fischer, R.; et al. Impaired autophagy induces chronic atrophic pancreatitis in mice via sex- and nutrition-dependent processes. Gastroenterology 2015, 148, 626–638.e17. [Google Scholar] [CrossRef] [PubMed]
  272. Perides, G.; Weiss, E.R.; Michael, E.S.; Laukkarinen, J.M.; Duffield, J.S.; Steer, M.L. TNF-alpha-dependent regulation of acute pancreatitis severity by Ly-6C(hi) monocytes in mice. J. Biol. Chem. 2011, 286, 13327–13335. [Google Scholar] [CrossRef] [PubMed]
  273. Fink, S.L.; Cookson, B.T. Caspase-1-dependent pore formation during pyroptosis leads to osmotic lysis of infected host macrophages. Cell Microbiol. 2006, 8, 1812–1825. [Google Scholar] [CrossRef] [PubMed]
  274. Bergsbaken, T.; Fink, S.L.; Cookson, B.T. Pyroptosis: Host cell death and inflammation. Nat. Rev. Microbiol. 2009, 7, 99–109. [Google Scholar] [CrossRef] [PubMed]
  275. Pan, B.; Li, Y.; Liu, Y.; Wang, W.; Huang, G.; Ouyang, Y. Circulating CitH3 Is a Reliable Diagnostic and Prognostic Biomarker of Septic Patients in Acute Pancreatitis. Front. Immunol. 2021, 12, 766391. [Google Scholar] [CrossRef] [PubMed]
  276. Merza, M.; Hartman, H.; Rahman, M.; Hwaiz, R.; Zhang, E.; Renström, E.; Luo, L.; Mörgelin, M.; Regner, S.; Thorlacius, H. Neutrophil Extracellular Traps Induce Trypsin Activation, Inflammation, and Tissue Damage in Mice With Severe Acute Pancreatitis. Gastroenterology 2015, 149, 1920–1931.e8. [Google Scholar] [CrossRef]
  277. Leppkes, M.; Schick, M.; Hohberger, B.; Mahajan, A.; Knopf, J.; Schett, G.; Muñoz, L.E.; Herrmann, M. Updates on NET formation in health and disease. Semin Arthritis Rheum. 2019, 49, S43–S48. [Google Scholar] [CrossRef] [PubMed]
  278. Cahilog, Z.; Zhao, H.; Wu, L.; Alam, A.; Eguchi, S.; Weng, H.; Ma, D. The Role of Neutrophil NETosis in Organ Injury: Novel Inflammatory Cell Death Mechanisms. Inflammation 2020, 43, 2021–2032. [Google Scholar] [CrossRef] [PubMed]
  279. Murthy, P.; Singhi, A.D.; Ross, M.A.; Loughran, P.; Paragomi, P.; Papachristou, G.I.; Whitcomb, D.C.; Zureikat, A.H.; Lotze, M.T.; Zeh Iii, H.J.; et al. Enhanced Neutrophil Extracellular Trap Formation in Acute Pancreatitis Contributes to Disease Severity and Is Reduced by Chloroquine. Front. Immunol. 2019, 10, 28. [Google Scholar] [CrossRef] [PubMed]
  280. Wan, J.; Ren, Y.; Yang, X.; Li, X.; Xia, L.; Lu, N. The Role of Neutrophils and Neutrophil Extracellular Traps in Acute Pancreatitis. Front. Cell Dev. Biol. 2021, 8, 565758. [Google Scholar] [CrossRef] [PubMed]
  281. Hu, J.; Kang, H.; Chen, H.; Yao, J.; Yi, X.; Tang, W.; Wan, M. Targeting neutrophil extracellular traps in severe acute pancreatitis treatment. Therap. Adv. Gastroenterol. 2020, 13, 1756284820974913. [Google Scholar] [CrossRef] [PubMed]
  282. Zhou, X.; Jin, S.; Pan, J.; Lin, Q.; Yang, S.; Ambe, P.C.; Basharat, Z.; Zimmer, V.; Wang, W.; Hong, W. Damage associated molecular patterns and neutrophil extracellular traps in acute pancreatitis. Front. Cell Infect. Microbiol. 2022, 12, 927193. [Google Scholar] [CrossRef] [PubMed]
  283. Mentula, P.; Kylänpää, M.L.; Kemppainen, E.; Jansson, S.E.; Sarna, S.; Puolakkainen, P.; Haapiainen, R.; Repo, H. Plasma anti-inflammatory cytokines and monocyte human leucocyte antigen-DR expression in patients with acute pancreatitis. Scand J. Gastroenterol. 2004, 39, 178–187. [Google Scholar] [CrossRef] [PubMed]
  284. Zhang, R.; Shi, J.; Zhang, R.; Ni, J.; Habtezion, A.; Wang, X.; Hu, G.; Xue, J. Expanded CD14hiCD16- Immunosuppressive Monocytes Predict Disease Severity in Patients with Acute Pancreatitis. J. Immunol. 2019, 202, 2578–2584. [Google Scholar] [CrossRef] [PubMed]
  285. Pan, T.; Zhou, T.; Li, L.; Liu, Z.; Chen, Y.; Mao, E.; Li, M.; Qu, H.; Liu, J. Monocyte programmed death ligand-1 expression is an early marker for predicting infectious complications in acute pancreatitis. Crit. Care 2017, 21, 186. [Google Scholar] [CrossRef] [PubMed]
  286. Liu, H.; Li, W.; Wang, X.; Li, J.; Yu, W. Early gut mucosal dysfunction in patients with acute pancreatitis. Pancreas 2008, 36, 192–196. [Google Scholar] [CrossRef] [PubMed]
  287. Rahman, S.H.; Ammori, B.J.; Holmfield, J.; Larvin, M.; McMahon, M.J. Intestinal hypoperfusion contributes to gut barrier failure in severe acute pancreatitis. J. Gastrointest. Surg. 2003, 7, 26–36. [Google Scholar] [CrossRef] [PubMed]
  288. Li, J.P.; Yang, J.; Huang, J.R.; Jiang, D.L.; Zhang, F.; Liu, M.F.; Qiang, Y.; Gu, Y.L. Immunosuppression and the infection caused by gut mucosal barrier dysfunction in patients with early severe acute pancreatitis. Front. Biosci. 2013, 18, 892–900. [Google Scholar] [CrossRef]
  289. Venkatesh, K.; Glenn, H.; Delaney, A.; Andersen, C.R.; Sasson, S.C. Fire in the belly: A scoping review of the immunopathological mechanisms of acute pancreatitis. Front. Immunol. 2023, 13, 1077414. [Google Scholar] [CrossRef] [PubMed]
  290. Bonjoch, L.; Casas, V.; Carrascal, M.; Closa, D. Involvement of exosomes in lung inflammation associated with experimental acute pancreatitis. J. Pathol. 2016, 240, 235–245. [Google Scholar] [CrossRef]
  291. Yang, Y.; Huang, Q.; Luo, C.; Wen, Y.; Liu, R.; Sun, H.; Tang, L. MicroRNAs in acute pancreatitis: From pathogenesis to novel diagnosis and therapy. J. Cell Physiol. 2020, 235, 1948–1961. [Google Scholar] [CrossRef] [PubMed]
  292. Munir, F.; Jamshed, M.B.; Shahid, N.; Muhammad, S.A.; Ghanem, N.B.; Qiyu, Z. Current status of diagnosis and Mesenchymal stem cells therapy for acute pancreatitis. Physiol. Rep. 2019, 7, e14170. [Google Scholar] [CrossRef] [PubMed]
  293. Hasan, A.; Moscoso, D.I.; Kastrinos, F. The Role of Genetics in Pancreatitis. Gastrointest. Endosc. Clin. N. Am. 2018, 28, 587–603. [Google Scholar] [CrossRef] [PubMed]
  294. Zator, Z.; Whitcomb, D.C. Insights into the genetic risk factors for the development of pancreatic disease. Therap. Adv. Gastroenterol. 2017, 10, 323–336. [Google Scholar] [CrossRef] [PubMed]
  295. van Geenen, E.J.; Smits, M.M.; Schreuder, T.C.; van der Peet, D.L.; Bloemena, E.; Mulder, C.J. Smoking is related to pancreatic fibrosis in humans. Am. J. Gastroenterol. 2011, 106, 1161–1166; quiz 1167. [Google Scholar] [CrossRef] [PubMed]
  296. Klöppel, G.; Maillet, B. Pseudocysts in chronic pancreatitis: A morphological analysis of 57 resection specimens and 9 autopsy pancreata. Pancreas 1991, 6, 266–274. [Google Scholar] [CrossRef] [PubMed]
  297. Klöppel, G.; Maillet, B. The morphological basis for the evolution of acute pancreatitis into chronic pancreatitis. Virchows Arch. A Pathol. Anat. Histopathol. 1992, 420, 1–4. [Google Scholar] [CrossRef] [PubMed]
  298. Bhanot, U.K.; Möller, P. Mechanisms of parenchymal injury and signaling pathways in ectatic ducts of chronic pancreatitis: Implications for pancreatic carcinogenesis. Lab. Investig. 2009, 89, 489–497. [Google Scholar] [CrossRef] [PubMed]
  299. Witt, H.; Apte, M.V.; Keim, V.; Wilson, J.S. Chronic pancreatitis: Challenges and advances in pathogenesis, genetics, diagnosis, and therapy. Gastroenterology 2007, 132, 1557–1573. [Google Scholar] [CrossRef] [PubMed]
  300. Yadav, D.; Whitcomb, D.C. The role of alcohol and smoking in pancreatitis. Nat. Rev. Gastroenterol. Hepatol. 2010, 7, 131–145. [Google Scholar] [CrossRef]
  301. Whitcomb, D.C. Hereditary pancreatitis: New insights into acute and chronic pancreatitis. Gut 1999, 45, 317–322. [Google Scholar] [CrossRef] [PubMed]
  302. Deng, X.; Wang, L.; Elm, M.S.; Gabazadeh, D.; Diorio, G.J.; Eagon, P.K.; Whitcomb, D.C. Chronic alcohol consumption accelerates fibrosis in response to cerulein-induced pancreatitis in rats. Am. J. Pathol. 2005, 166, 93–106. [Google Scholar] [CrossRef] [PubMed]
  303. Leung, P.S.; Chan, Y.C. Role of oxidative stress in pancreatic inflammation. Antioxid. Redox Signal. 2009, 11, 135–165. [Google Scholar] [CrossRef] [PubMed]
  304. Saito, I.; Hashimoto, S.; Saluja, A.; Steer, M.L.; Meldolesi, J. Intracellular transport of pancreatic zymogens during caerulein supramaximal stimulation. Am. J. Physiol. 1987, 253 Pt 1, G517–G526. [Google Scholar] [CrossRef] [PubMed]
  305. Norton, I.D.; Apte, M.V.; Lux, O.; Haber, P.S.; Pirola, R.C.; Wilson, J.S. Chronic ethanol administration causes oxidative stress in the rat pancreas. J. Lab. Clin. Med. 1998, 131, 442–446. [Google Scholar] [CrossRef] [PubMed]
  306. Yamaguchi, M.; Steward, M.C.; Smallbone, K.; Sohma, Y.; Yamamoto, A.; Ko, S.B.; Kondo, T.; Ishiguro, H. Bicarbonate-rich fluid secretion predicted by a computational model of guinea-pig pancreatic duct epithelium. J. Physiol. 2017, 595, 1947–1972. [Google Scholar] [CrossRef] [PubMed]
  307. Hall, P.A.; Lemoine, N.R. Rapid acinar to ductal transdifferentiation in cultured human exocrine pancreas. J. Pathol. 1992, 166, 97–103. [Google Scholar] [CrossRef]
  308. Houbracken, I.; de Waele, E.; Lardon, J.; Ling, Z.; Heimberg, H.; Rooman, I.; Bouwens, L. Lineage tracing evidence for transdifferentiation of acinar to duct cells and plasticity of human pancreas. Gastroenterology 2011, 141, 731–741.e4. [Google Scholar] [CrossRef]
  309. Weiss, F.U.; Skube, M.E.; Lerch, M.M. Chronic pancreatitis: An update on genetic risk factors. Curr. Opin. Gastroenterol. 2018, 34, 322–329. [Google Scholar] [CrossRef] [PubMed]
  310. Moore, P.C.; Cortez, J.T.; Chamberlain, C.E.; Alba, D.; Berger, A.C.; Quandt, Z.; Chan, A.; Cheng, M.H.; Bautista, J.L.; Peng, J.; et al. Elastase 3B mutation links to familial pancreatitis with diabetes and pancreatic adenocarcinoma. J. Clin. Investig. 2019, 129, 4676–4681. [Google Scholar] [CrossRef] [PubMed]
  311. Rebours, V.; Vullierme, M.P.; Hentic, O.; Maire, F.; Hammel, P.; Ruszniewski, P.; Lévy, P. Smoking and the course of recurrent acute and chronic alcoholic pancreatitis: A dose-dependent relationship. Pancreas 2012, 41, 1219–1224. [Google Scholar] [CrossRef] [PubMed]
  312. Nikkola, J.; Räty, S.; Laukkarinen, J.; Seppänen, H.; Lappalainen-Lehto, R.; Järvinen, S.; Nordback, I.; Sand, J. Abstinence after first acute alcohol-associated pancreatitis protects against recurrent pancreatitis and minimizes the risk of pancreatic dysfunction. Alcohol. Alcohol. 2013, 48, 483–486. [Google Scholar] [CrossRef] [PubMed]
  313. Werner, J.; Laposata, M.; Fernández-del Castillo, C.; Saghir, M.; Iozzo, R.V.; Lewandrowski, K.B.; Warshaw, A.L. Pancreatic injury in rats induced by fatty acid ethyl ester, a nonoxidative metabolite of alcohol. Gastroenterology 1997, 113, 286–294. [Google Scholar] [CrossRef]
  314. Vonlaufen, A.; Wilson, J.S.; Pirola, R.C.; Apte, M.V. Role of alcohol metabolism in chronic pancreatitis. Alcohol. Res. Health. 2007, 30, 48–54. [Google Scholar] [PubMed]
  315. Gu, H.; Werner, J.; Bergmann, F.; Whitcomb, D.C.; Büchler, M.W.; Fortunato, F. Necro-inflammatory response of pancreatic acinar cells in the pathogenesis of acute alcoholic pancreatitis. Cell Death Dis. 2013, 4, e816. [Google Scholar] [CrossRef] [PubMed]
  316. Apte, M.V.; Wilson, J.S. Stellate cell activation in alcoholic pancreatitis. Pancreas 2003, 27, 316–320. [Google Scholar] [CrossRef] [PubMed]
  317. Hu, F.; Lou, N.; Jiao, J.; Guo, F.; Xiang, H.; Shang, D. Macrophages in pancreatitis: Mechanisms and therapeutic potential. Biomed. Pharmacother. 2020, 131, 110693. [Google Scholar] [CrossRef]
  318. Detlefsen, S.; Sipos, B.; Feyerabend, B.; Klöppel, G. Fibrogenesis in alcoholic chronic pancreatitis: The role of tissue necrosis, macrophages, myofibroblasts and cytokines. Mod. Pathol. 2006, 19, 1019–1026. [Google Scholar] [CrossRef] [PubMed]
  319. Shek, F.W.; Benyon, R.C.; Walker, F.M.; McCrudden, P.R.; Pender, S.L.; Williams, E.J.; Johnson, P.A.; Johnson, C.D.; Bateman, A.C.; Fine, D.R.; et al. Expression of transforming growth factor-beta 1 by pancreatic stellate cells and its implications for matrix secretion and turnover in chronic pancreatitis. Am. J. Pathol. 2002, 160, 1787–1798. [Google Scholar] [CrossRef] [PubMed]
  320. Bynigeri, R.R.; Jakkampudi, A.; Jangala, R.; Subramanyam, C.; Sasikala, M.; Rao, G.V.; Reddy, D.N.; Talukdar, R. Pancreatic stellate cell: Pandora’s box for pancreatic disease biology. World J. Gastroenterol. 2017, 23, 382–405. [Google Scholar] [CrossRef] [PubMed]
  321. Zhang, H.; Liu, B.; Xu, X.F.; Jiang, T.T.; Zhang, X.Q.; Shi, Y.L.; Chen, Y.; Liu, F.; Gu, J.; Zhu, L.J.; et al. Pathophysiology of chronic pancreatitis induced by dibutyltin dichloride joint ethanol in mice. World J. Gastroenterol. 2016, 22, 2960–2970. [Google Scholar] [CrossRef] [PubMed]
  322. Wu, J.; Zhang, L.; Shi, J.; He, R.; Yang, W.; Habtezion, A.; Niu, N.; Lu, P.; Xue, J. Macrophage phenotypic switch orchestrates the inflammation and repair/regeneration following acute pancreatitis injury. eBioMedicine 2020, 58, 102920. [Google Scholar] [CrossRef] [PubMed]
  323. Shapouri-Moghaddam, A.; Mohammadian, S.; Vazini, H.; Taghadosi, M.; Esmaeili, S.A.; Mardani, F.; Seifi, B.; Mohammadi, A.; Afshari, J.T.; Sahebkar, A. Macrophage plasticity, polarization, and function in health and disease. J. Cell Physiol. 2018, 233, 6425–6440. [Google Scholar] [CrossRef] [PubMed]
  324. Kishore, A.; Petrek, M. Roles of Macrophage Polarization and Macrophage-Derived miRNAs in Pulmonary Fibrosis. Front. Immunol. 2021, 12, 678457. [Google Scholar] [CrossRef] [PubMed]
  325. Lu, Y.; Lu, G.; Gao, L.; Zhu, Q.; Xue, J.; Zhang, J.; Ma, X.; Ma, N.; Yang, Q.; Dong, J.; et al. The Proresolving Lipid Mediator Maresin1 Alleviates Experimental Pancreatitis via Switching Macrophage Polarization. Mediat. Inflamm. 2021, 2021, 6680456. [Google Scholar] [CrossRef] [PubMed]
  326. Alho, H.; Sillanaukee, P.; Kalela, A.; Jaakkola, O.; Laine, S.; Nikkari, S.T. Alcohol misuse increases serum antibodies to oxidized LDL and C-reactive protein. Alcohol. Alcohol. 2004, 39, 312–315. [Google Scholar] [CrossRef] [PubMed]
  327. Gonzalez-Quintela, A.; Campos, J.; Loidi, L.; Quinteiro, C.; Perez, L.F.; Gude, F. Serum TNF-alpha levels in relation to alcohol consumption and common TNF gene polymorphisms. Alcohol 2008, 42, 513–518. [Google Scholar] [CrossRef] [PubMed]
  328. Apte, M.V.; Pirola, R.C.; Wilson, J.S. Mechanisms of alcoholic pancreatitis. J. Gastroenterol. Hepatol. 2010, 25, 1816–1826. [Google Scholar] [CrossRef]
  329. Xue, J.; Sharma, V.; Hsieh, M.H.; Chawla, A.; Murali, R.; Pandol, S.J.; Habtezion, A. Alternatively activated macrophages promote pancreatic fibrosis in chronic pancreatitis. Nat. Commun. 2015, 6, 7158. [Google Scholar] [CrossRef] [PubMed]
  330. Żorniak, M.; Sirtl, S.; Mayerle, J.; Beyer, G. What Do We Currently Know about the Pathophysiology of Alcoholic Pancreatitis: A Brief Review. Visc. Med. 2020, 36, 182–190. [Google Scholar] [CrossRef]
  331. Vogelmann, R.; Ruf, D.; Wagner, M.; Adler, G.; Menke, A. Effects of fibrogenic mediators on the development of pancreatic fibrosis in a TGF-beta1 transgenic mouse model. Am. J. Physiol. Gastrointest. Liver Physiol. 2001, 280, G164–G172. [Google Scholar] [CrossRef]
  332. Manohar, M.; Verma, A.K.; Venkateshaiah, S.U.; Sanders, N.L.; Mishra, A. Pathogenic mechanisms of pancreatitis. World J. Gastrointest. Pharmacol. Ther. 2017, 8, 10–25. [Google Scholar] [CrossRef] [PubMed]
  333. Tang, D.; Wu, Q.; Zhang, J.; Zhang, H.; Yuan, Z.; Xu, J.; Chong, Y.; Huang, Y.; Xiong, Q.; Wang, S.; et al. Galectin-1 expression in activated pancreatic satellite cells promotes fibrosis in chronic pancreatitis/pancreatic cancer via the TGF-β1/Smad pathway. Oncol. Rep. 2018, 39, 1347–1355. [Google Scholar] [CrossRef] [PubMed]
  334. Lin, H.; Dong, B.; Qi, L.; Wei, Y.; Zhang, Y.; Cai, X.; Zhang, Q.; Li, J.; Li, L. Inhibitory Smads suppress pancreatic stellate cell activation through negative feedback in chronic pancreatitis. Ann. Transl. Med. 2021, 9, 384. [Google Scholar] [CrossRef] [PubMed]
  335. An, W.; Zhu, J.W.; Jiang, F.; Jiang, H.; Zhao, J.L.; Liu, M.Y.; Li, G.X.; Shi, X.G.; Sun, C.; Li, Z.S. Fibromodulin is upregulated by oxidative stress through the MAPK/AP-1 pathway to promote pancreatic stellate cell activation. Pancreatology 2020, 20, 278–287. [Google Scholar] [CrossRef] [PubMed]
  336. Xu, X.F.; Liu, F.; Xin, J.Q.; Fan, J.W.; Wu, N.; Zhu, L.J.; Duan, L.F.; Li, Y.Y.; Zhang, H. Respective roles of the mitogen-activated protein kinase (MAPK) family members in pancreatic stellate cell activation induced by transforming growth factor-β1 (TGF-β1). Biochem. Biophys. Res. Commun. 2018, 501, 365–373. [Google Scholar] [CrossRef] [PubMed]
  337. Jin, G.; Hong, W.; Guo, Y.; Bai, Y.; Chen, B. Molecular Mechanism of Pancreatic Stellate Cells Activation in Chronic Pancreatitis and Pancreatic Cancer. J. Cancer. 2020, 11, 1505–1515. [Google Scholar] [CrossRef] [PubMed]
  338. Ramakrishnan, P.; Loh, W.M.; Gopinath, S.C.B.; Bonam, S.R.; Fareez, I.M.; Mac Guad, R.; Sim, M.S.; Wu, Y.S. Selective phytochemicals targeting pancreatic stellate cells as new anti-fibrotic agents for chronic pancreatitis and pancreatic cancer. Acta Pharm. Sin B 2020, 10, 399–413. [Google Scholar] [CrossRef] [PubMed]
  339. Xue, R.; Jia, K.; Wang, J.; Yang, L.; Wang, Y.; Gao, L.; Hao, J. A Rising Star in Pancreatic Diseases: Pancreatic Stellate Cells. Front. Physiol. 2018, 9, 754. [Google Scholar] [CrossRef] [PubMed]
  340. Shimosegawa, T. A New Insight into Chronic Pancreatitis. Tohoku J. Exp. Med. 2019, 248, 225–238. [Google Scholar] [CrossRef] [PubMed]
  341. Luttenberger, T.; Schmid-Kotsas, A.; Menke, A.; Siech, M.; Beger, H.; Adler, G.; Grünert, A.; Bachem, M.G. Platelet-derived growth factors stimulate proliferation and extracellular matrix synthesis of pancreatic stellate cells: Implications in pathogenesis of pancreas fibrosis. Lab. Investig. 2000, 80, 47–55. [Google Scholar] [CrossRef] [PubMed]
  342. Schneider, E.; Schmid-Kotsas, A.; Zhao, J.; Weidenbach, H.; Schmid, R.M.; Menke, A.; Adler, G.; Waltenberger, J.; Grünert, A.; Bachem, M.G. Identification of mediators stimulating proliferation and matrix synthesis of rat pancreatic stellate cells. Am. J. Physiol. Cell Physiol. 2001, 281, C532–C543. [Google Scholar] [CrossRef] [PubMed]
  343. Phillips, P.A.; Wu, M.J.; Kumar, R.K.; Doherty, E.; McCarroll, J.A.; Park, S.; Pirola, R.C.; Wilson, J.S.; Apte, M.V. Cell migration: A novel aspect of pancreatic stellate cell biology. Gut 2003, 52, 677–682. [Google Scholar] [CrossRef]
  344. di Mola, F.F.; Friess, H.; Martignoni, M.E.; Di Sebastiano, P.; Zimmermann, A.; Innocenti, P.; Graber, H.; Gold, L.I.; Korc, M.; Büchler, M.W. Connective tissue growth factor is a regulator for fibrosis in human chronic pancreatitis. Ann. Surg. 1999, 230, 63–71. [Google Scholar] [CrossRef] [PubMed]
  345. Gao, R.; Brigstock, D.R. Connective tissue growth factor (CCN2) in rat pancreatic stellate cell function: Integrin alpha5beta1 as a novel CCN2 receptor. Gastroenterology 2005, 129, 1019–1030. [Google Scholar] [CrossRef] [PubMed]
  346. Karger, A.; Fitzner, B.; Brock, P.; Sparmann, G.; Emmrich, J.; Liebe, S.; Jaster, R. Molecular insights into connective tissue growth factor action in rat pancreatic stellate cells. Cell Signal. 2008, 20, 1865–1872. [Google Scholar] [CrossRef] [PubMed]
  347. Mews, P.; Phillips, P.; Fahmy, R.; Korsten, M.; Pirola, R.; Wilson, J.; Apte, M. Pancreatic stellate cells respond to inflammatory cytokines: Potential role in chronic pancreatitis. Gut 2002, 50, 535–541. [Google Scholar] [CrossRef] [PubMed]
  348. Marzoq, A.J.; Giese, N.; Hoheisel, J.D.; Alhamdani, M.S.S. Proteome variations in pancreatic stellate cells upon stimulation with proinflammatory factors. J. Biol. Chem. 2013, 288, 32517–32527. [Google Scholar] [CrossRef] [PubMed]
  349. Charo, C.; Holla, V.; Arumugam, T.; Hwang, R.; Yang, P.; Dubois, R.N.; Menter, D.G.; Logsdon, C.D.; Ramachandran, V. Prostaglandin E2 regulates pancreatic stellate cell activity via the EP4 receptor. Pancreas 2013, 42, 467–474. [Google Scholar] [CrossRef] [PubMed]
  350. Huang, H.; Chen, J.; Peng, L.; Yao, Y.; Deng, D.; Zhang, Y.; Liu, Y.; Wang, H.; Li, Z.; Bi, Y.; et al. Transgenic expression of cyclooxygenase-2 in pancreatic acinar cells induces chronic pancreatitis. Am. J. Physiol. Gastrointest. Liver Physiol. 2019, 316, G179–G186. [Google Scholar] [CrossRef] [PubMed]
  351. Masamune, A.; Kikuta, K.; Watanabe, T.; Satoh, K.; Satoh, A.; Shimosegawa, T. Pancreatic stellate cells express Toll-like receptors. J. Gastroenterol. 2008, 43, 352–362. [Google Scholar] [CrossRef] [PubMed]
  352. Xue, J.; Zhao, Q.; Sharma, V.; Nguyen, L.P.; Lee, Y.N.; Pham, K.L.; Edderkaoui, M.; Pandol, S.J.; Park, W.; Habtezion, A. Aryl Hydrocarbon Receptor Ligands in Cigarette Smoke Induce Production of Interleukin-22 to Promote Pancreatic Fibrosis in Models of Chronic Pancreatitis. Gastroenterology 2016, 151, 1206–1217. [Google Scholar] [CrossRef] [PubMed]
  353. Lee, A.T.; Xu, Z.; Pothula, S.P.; Patel, M.B.; Pirola, R.C.; Wilson, J.S.; Apte, M.V. Alcohol and cigarette smoke components activate human pancreatic stellate cells: Implications for the progression of chronic pancreatitis. Alcohol. Clin. Exp. Res. 2015, 39, 2123–2133. [Google Scholar] [CrossRef] [PubMed]
  354. Gukovsky, I.; Pandol, S.J.; Mareninova, O.A.; Shalbueva, N.; Jia, W.; Gukovskaya, A.S. Impaired autophagy and organellar dysfunction in pancreatitis. J. Gastroenterol. Hepatol. 2012, 27 (Suppl. S2), 27–32. [Google Scholar] [CrossRef] [PubMed]
  355. Li, S.; Xie, K. Ductal metaplasia in pancreas. Biochim. Biophys. Acta Rev. Cancer 2022, 1877, 188698. [Google Scholar] [CrossRef] [PubMed]
  356. Parte, S.; Nimmakayala, R.K.; Batra, S.K.; Ponnusamy, M.P. Acinar to ductal cell trans-differentiation: A prelude to dysplasia and pancreatic ductal adenocarcinoma. Biochim. Biophys. Acta Rev. Cancer 2022, 1877, 188669. [Google Scholar] [CrossRef] [PubMed]
  357. Masamune, A.; Kikuta, K.; Watanabe, T.; Satoh, K.; Hirota, M.; Shimosegawa, T. Hypoxia stimulates pancreatic stellate cells to induce fibrosis and angiogenesis in pancreatic cancer. Am. J. Physiol. Gastrointest. Liver Physiol. 2008, 295, G709–G717. [Google Scholar] [CrossRef] [PubMed]
  358. Swain, S.M.; Romac, J.M.; Vigna, S.R.; Liddle, R.A. Piezo1-mediated stellate cell activation causes pressure-induced pancreatic fibrosis in mice. JCI Insight 2022, 7, e158288. [Google Scholar] [CrossRef] [PubMed]
  359. Cannon, A.; Thompson, C.M.; Bhatia, R.; Armstrong, K.A.; Solheim, J.C.; Kumar, S.; Batra, S.K. Molecular mechanisms of pancreatic myofibroblast activation in chronic pancreatitis and pancreatic ductal adenocarcinoma. J. Gastroenterol. 2021, 56, 689–703. [Google Scholar] [CrossRef] [PubMed]
  360. Li, B.Q.; Liu, X.Y.; Mao, T.; Zheng, T.H.; Zhang, P.; Zhang, Q.; Zhang, Y.; Li, X.Y. The research progress of anti-inflammatory and anti-fibrosis treatment of chronic pancreatitis. Front. Oncol. 2022, 12, 1050274. [Google Scholar] [CrossRef] [PubMed]
  361. Kessler, A.; Weksler-Zangen, S.; Ilan, Y. Role of the Immune System and the Circadian Rhythm in the Pathogenesis of Chronic Pancreatitis: Establishing a Personalized Signature for Improving the Effect of Immunotherapies for Chronic Pancreatitis. Pancreas 2020, 49, 1024–1032. [Google Scholar] [CrossRef] [PubMed]
  362. Hunger, R.E.; Mueller, C.; Z’graggen, K.; Friess, H.; Büchler, M.W. Cytotoxic cells are activated in cellular infiltrates of alcoholic chronic pancreatitis. Gastroenterology 1997, 112, 1656–1663. [Google Scholar] [CrossRef] [PubMed]
  363. Schmitz-Winnenthal, H.; Pietsch, D.H.; Schimmack, S.; Bonertz, A.; Udonta, F.; Ge, Y.; Galindo, L.; Specht, S.; Volk, C.; Zgraggen, K.; et al. Chronic pancreatitis is associated with disease-specific regulatory T-cell responses. Gastroenterology 2010, 138, 1178–1188. [Google Scholar] [CrossRef] [PubMed]
  364. Grundsten, M.; Liu, G.Z.; Permert, J.; Hjelmstrom, P.; Tsai, J.A. Increased central memory T cells in patients with chronic pancreatitis. Pancreatology 2005, 5, 177–182. [Google Scholar] [CrossRef] [PubMed]
  365. Kist, M.; Vucic, D. Cell death pathways: Intricate connections and disease implications. EMBO J. 2021, 40, e106700. [Google Scholar] [CrossRef] [PubMed]
  366. Tang, D.; Kang, R.; Berghe, T.V.; Vandenabeele, P.; Kroemer, G. The molecular machinery of regulated cell death. Cell Res. 2019, 29, 347–364. [Google Scholar] [CrossRef] [PubMed]
  367. Lee, P.J.; Papachristou, G.I. New insights into acute pancreatitis. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 479–496. [Google Scholar] [CrossRef]
  368. Han, X.; Li, B.; Bao, J.; Wu, Z.; Chen, C.; Ni, J.; Shen, J.; Song, P.; Peng, Q.; Wan, R.; et al. Endoplasmic reticulum stress promoted acinar cell necroptosis in acute pancreatitis through cathepsinB-mediated AP-1 activation. Front. Immunol. 2022, 13, 968639. [Google Scholar] [CrossRef] [PubMed]
  369. Ma, N.; Yuan, C.; Shi, J.; Zhu, Q.; Liu, Y.; Ma, X.; Li, B.; Gong, W.; Xue, J.; Lu, G.; et al. Interleukin-37 protects against acinar cell pyroptosis in acute pancreatitis. JCI Insight. 2022, 7, e161244. [Google Scholar] [CrossRef]
  370. Li, H.Y.; Lin, Y.J.; Zhang, L.; Zhao, J.; Xiao, D.Y.; Huang, Z.Z.; Li, P.W. Progress of pyroptosis in acute pancreatitis. Chin. Med. J. 2021, 134, 2160–2162. [Google Scholar] [CrossRef]
  371. Chen, W.; Yuan, C.; Lu, Y.; Zhu, Q.; Ma, X.; Xiao, W.; Gong, W.; Huang, W.; Xia, Q.; Lu, G.; et al. Tanshinone IIA Protects against Acute Pancreatitis in Mice by Inhibiting Oxidative Stress via the Nrf2/ROS Pathway. Oxid. Med. Cell Longev. 2020, 2020, 5390482. [Google Scholar] [CrossRef] [PubMed]
  372. Zhou, X.; Fu, Y.; Liu, W.; Mu, Y.; Zhang, H.; Chen, J.; Liu, P. Ferroptosis in Chronic Liver Diseases: Opportunities and Challenges. Front. Mol. Biosci. 2022, 9, 928321. [Google Scholar] [CrossRef] [PubMed]
  373. Meng, Y.T.; Zhou, Y.; Han, P.Y.; Ren, H.B. Ferroptosis inhibition attenuates inflammatory response in mice with acute hypertriglyceridemic pancreatitis. World J. Gastroenterol. 2023, 29, 2294–2309. [Google Scholar] [CrossRef] [PubMed]
  374. Fortunato, F.; Bürgers, H.; Bergmann, F.; Rieger, P.; Büchler, M.W.; Kroemer, G.; Werner, J. Impaired autolysosome formation correlates with Lamp-2 depletion: Role of apoptosis, autophagy, and necrosis in pancreatitis. Gastroenterology 2009, 137, 350–360.e5. [Google Scholar] [CrossRef] [PubMed]
  375. Li, N.; Wu, X.; Holzer, R.G.; Lee, J.H.; Todoric, J.; Park, E.J.; Ogata, H.; Gukovskaya, A.S.; Gukovsky, I.; Pizzo, D.P.; et al. Loss of acinar cell IKKα triggers spontaneous pancreatitis in mice. J. Clin. Investig. 2013, 123, 2231–2243. [Google Scholar] [CrossRef]
  376. Mareninova, O.A.; Hermann, K.; French, S.W.; O’Konski, M.S.; Pandol, S.J.; Webster, P.; Erickson, A.H.; Katunuma, N.; Gorelick, F.S.; Gukovsky, I.; et al. Impaired autophagic flux mediates acinar cell vacuole formation and trypsinogen activation in rodent models of acute pancreatitis. J. Clin. Investig. 2009, 119, 3340–3355. [Google Scholar] [CrossRef] [PubMed]
  377. Feng, D.; Park, O.; Radaeva, S.; Wang, H.; Yin, S.; Kong, X.; Zheng, M.; Zakhari, S.; Kolls, J.K.; Gao, B. Interleukin-22 ameliorates cerulein-induced pancreatitis in mice by inhibiting the autophagic pathway. Int. J. Biol. Sci. 2012, 8, 249–257. [Google Scholar] [CrossRef] [PubMed]
  378. Wu, Y.; Tang, L.; Wang, B.; Sun, Q.; Zhao, P.; Li, W. The role of autophagy in maintaining intestinal mucosal barrier. J. Cell Physiol. 2019, 234, 19406–19419. [Google Scholar] [CrossRef] [PubMed]
  379. Larabi, A.; Barnich, N.; Nguyen, H.T.T. New insights into the interplay between autophagy, gut microbiota and inflammatory responses in IBD. Autophagy 2020, 16, 38–51. [Google Scholar] [CrossRef] [PubMed]
  380. Wang, H.; Li, C.; Jiang, Y.; Li, H.; Zhang, D. Effects of Bacterial Translocation and Autophagy on Acute Lung Injury Induced by Severe Acute Pancreatitis. Gastroenterol. Res. Pract. 2020, 2020, 8953453. [Google Scholar] [CrossRef] [PubMed]
  381. Kong, L.; Deng, J.; Zhou, X.; Cai, B.; Zhang, B.; Chen, X.; Chen, Z.; Wang, W. Sitagliptin activates the p62-Keap1-Nrf2 signalling pathway to alleviate oxidative stress and excessive autophagy in severe acute pancreatitis-related acute lung injury. Cell Death Dis. 2021, 12, 928. [Google Scholar] [CrossRef] [PubMed]
  382. Dolai, S.; Liang, T.; Orabi, A.I.; Holmyard, D.; Xie, L.; Greitzer-Antes, D.; Kang, Y.; Xie, H.; Javed, T.A.; Lam, P.P.; et al. Pancreatitis-Induced Depletion of Syntaxin 2 Promotes Autophagy and Increases Basolateral Exocytosis. Gastroenterology 2018, 154, 1805–1821.e5. [Google Scholar] [CrossRef] [PubMed]
  383. Dolai, S.; Takahashi, T.; Qin, T.; Liang, T.; Xie, L.; Kang, F.; Miao, Y.F.; Xie, H.; Kang, Y.; Manuel, J.; et al. Pancreas-specific SNAP23 depletion prevents pancreatitis by attenuating pathological basolateral exocytosis and formation of trypsin-activating autolysosomes. Autophagy 2021, 17, 3068–3081. [Google Scholar] [CrossRef]
  384. Huangfu, Y.; Yu, X.; Wan, C.; Zhu, Y.; Wei, Z.; Li, F.; Wang, Y.; Zhang, K.; Li, S.; Dong, Y.; et al. Xanthohumol alleviates oxidative stress and impaired autophagy in experimental severe acute pancreatitis through inhibition of AKT/mTOR. Front. Pharmacol. 2023, 14, 1105726. [Google Scholar] [CrossRef]
  385. Inman, K.S.; Liu, Y.; Scotti Buzhardt, M.L.; Leitges, M.; Krishna, M.; Crawford, H.C.; Fields, A.P.; Murray, N.R. Prkci Regulates Autophagy and Pancreatic Tumorigenesis in Mice. Cancers 2022, 14, 796. [Google Scholar] [CrossRef] [PubMed]
  386. Massafra, V.; van Mil, S.W.C. Farnesoid X receptor: A “homeostat” for hepatic nutrient metabolism. Biochim. Biophys. Acta Mol. Basis Dis. 2018, 1864, 45–59. [Google Scholar] [CrossRef]
  387. Ding, L.; Yang, L.; Wang, Z.; Huang, W. Bile acid nuclear receptor FXR and digestive system diseases. Acta Pharm. Sin. B 2015, 5, 135–144. [Google Scholar] [CrossRef] [PubMed]
  388. Hao, H.; Cao, L.; Jiang, C.; Che, Y.; Zhang, S.; Takahashi, S.; Wang, G.; Gonzalez, F.J. Farnesoid X Receptor Regulation of the NLRP3 Inflammasome Underlies Cholestasis-Associated Sepsis. Cell Metab. 2017, 25, 856–867.e5. [Google Scholar] [CrossRef]
  389. Zheng, Y.; Sun, W.; Wang, Z.; Liu, J.; Shan, C.; He, C.; Li, B.; Hu, X.; Zhu, W.; Liu, L.; et al. Activation of Pancreatic Acinar FXR Protects against Pancreatitis via Osgin1-Mediated Restoration of Efficient Autophagy. Research 2022, 2022, 9784081. [Google Scholar] [CrossRef]
  390. Vaccaro, M.I. Zymophagy: Selective autophagy of secretory granules. Int. J. Cell Biol. 2012, 2012, 396705. [Google Scholar] [CrossRef]
  391. Grasso, D.; Ropolo, A.; Lo Ré, A.; Boggio, V.; Molejón, M.I.; Iovanna, J.L.; Gonzalez, C.D.; Urrutia, R.; Vaccaro, M.I. Zymophagy, a novel selective autophagy pathway mediated by VMP1-USP9x-p62, prevents pancreatic cell death. J. Biol. Chem. 2011, 286, 8308–8324. [Google Scholar] [CrossRef]
  392. Mareninova, O.A.; Dillon, D.L.; Wightman, C.J.M.; Yakubov, I.; Takahashi, T.; Gaisano, H.Y.; Munson, K.; Ohmuraya, M.; Dawson, D.; Gukovsky, I.; et al. Rab9 Mediates Pancreatic Autophagy Switch From Canonical to Noncanonical, Aggravating Experimental Pancreatitis. Cell Mol. Gastroenterol. Hepatol. 2022, 13, 599–622. [Google Scholar] [CrossRef] [PubMed]
  393. Hashimoto, D.; Ohmuraya, M.; Hirota, M.; Yamamoto, A.; Suyama, K.; Ida, S.; Okumura, Y.; Takahashi, E.; Kido, H.; Araki, K.; et al. Involvement of autophagy in trypsinogen activation within the pancreatic acinar cells. J. Cell Biol. 2008, 181, 1065–1072. [Google Scholar] [CrossRef]
  394. Ohmuraya, M.; Yamamura, K. Autophagy and acute pancreatitis: A novel autophagy theory for trypsinogen activation. Autophagy 2008, 4, 1060–1062. [Google Scholar] [CrossRef] [PubMed]
  395. Gukovskaya, A.S.; Gukovsky, I. Autophagy and pancreatitis. Am. J. Physiol. Gastrointest. Liver Physiol. 2012, 303, G993–G1003. [Google Scholar] [CrossRef] [PubMed]
  396. Voronina, S.; Chvanov, M.; De Faveri, F.; Mayer, U.; Wileman, T.; Criddle, D.; Tepikin, A. Autophagy, Acute Pancreatitis and the Metamorphoses of a Trypsinogen-Activating Organelle. Cells. 2022, 11, 2514. [Google Scholar] [CrossRef] [PubMed]
  397. Gukovskaya, A.S.; Gorelick, F.S.; Groblewski, G.E.; Mareninova, O.A.; Lugea, A.; Antonucci, L.; Waldron, R.T.; Habtezion, A.; Karin, M.; Pandol, S.J.; et al. Recent Insights Into the Pathogenic Mechanism of Pancreatitis: Role of Acinar Cell Organelle Disorders. Pancreas 2019, 48, 459–470. [Google Scholar] [CrossRef] [PubMed]
  398. Courreges, A.P.; Najenson, A.C.; Vatta, M.S.; Bianciotti, L.G. Atrial natriuretic peptide attenuates endoplasmic reticulum stress in experimental acute pancreatitis. Biochim. Biophys. Acta Mol. Basis Dis. 2019, 1865, 485–493. [Google Scholar] [CrossRef] [PubMed]
  399. Logsdon, C.D.; Ji, B. The role of protein synthesis and digestive enzymes in acinar cell injury. Nat. Rev. Gastroenterol. Hepatol. 2013, 10, 362–370. [Google Scholar] [CrossRef] [PubMed]
  400. Park, S.; Zuber, C.; Roth, J. Selective autophagy of cytosolic protein aggregates involves ribosome-free rough endoplasmic reticulum. Histochem. Cell Biol. 2020, 153, 89–99. [Google Scholar] [CrossRef] [PubMed]
  401. Calvo-Garrido, J.; Escalante, R. Autophagy dysfunction and ubiquitin-positive protein aggregates in Dictyostelium cells lacking Vmp1. Autophagy 2010, 6, 100–109. [Google Scholar] [CrossRef] [PubMed]
  402. Xu, B.; Bai, B.; Sha, S.; Yu, P.; An, Y.; Wang, S.; Kong, X.; Liu, C.; Wei, N.; Feng, Q.; et al. Interleukin-1β induces autophagy by affecting calcium homeostasis and trypsinogen activation in pancreatic acinar cells. Int. J. Clin. Exp. Pathol. 2014, 7, 3620–3631. [Google Scholar] [PubMed]
  403. Dolai, S.; Liang, T.; Orabi, A.I.; Xie, L.; Holmyard, D.; Javed, T.A.; Fernandez, N.A.; Xie, H.; Cattral, M.S.; Thurmond, D.C.; et al. Depletion of the membrane-fusion regulator Munc18c attenuates caerulein hyperstimulation-induced pancreatitis. J. Biol. Chem. 2018, 293, 2510–2522. [Google Scholar] [CrossRef] [PubMed]
  404. Walter, P.; Ron, D. The unfolded protein response: From stress pathway to homeostatic regulation. Science 2011, 334, 1081–1086. [Google Scholar] [CrossRef] [PubMed]
  405. Lugea, A.; Waldron, R.T.; French, S.W.; Pandol, S.J. Drinking and driving pancreatitis: Links between endoplasmic reticulum stress and autophagy. Autophagy 2011, 7, 783–785. [Google Scholar] [CrossRef] [PubMed]
  406. Hall, J.C.; Crawford, H.C. The conspiracy of autophagy, stress and inflammation in acute pancreatitis. Curr. Opin. Gastroenterol. 2014, 30, 495–499. [Google Scholar] [CrossRef]
  407. Mareninova, O.A.; Jia, W.; Gretler, S.R.; Holthaus, C.L.; Thomas, D.D.H.; Pimienta, M.; Dillon, D.L.; Gukovskaya, A.S.; Gukovsky, I.; Groblewski, G.E. Transgenic expression of GFP-LC3 perturbs autophagy in exocrine pancreas and acute pancreatitis responses in mice. Autophagy 2020, 16, 2084–2097. [Google Scholar] [CrossRef] [PubMed]
  408. Chen, Y.; Zhang, J.; Zhao, Q.; Chen, Q.; Sun, Y.; Jin, Y.; Wu, J. Melatonin Induces Anti-Inflammatory Effects to Play a Protective Role via Endoplasmic Reticulum Stress in Acute Pancreatitis. Cell Physiol. Biochem. 2016, 40, 1094–1104. [Google Scholar] [CrossRef] [PubMed]
  409. Fazio, E.N.; Dimattia, G.E.; Chadi, S.A.; Kernohan, K.D.; Pin, C.L. Stanniocalcin 2 alters PERK signalling and reduces cellular injury during cerulein induced pancreatitis in mice. BMC Cell Biol. 2011, 12, 17. [Google Scholar] [CrossRef] [PubMed]
  410. Biczó, G.; Hegyi, P.; Dósa, S.; Shalbuyeva, N.; Berczi, S.; Sinervirta, R.; Hracskó, Z.; Siska, A.; Kukor, Z.; Jármay, K.; et al. The crucial role of early mitochondrial injury in L-lysine-induced acute pancreatitis. Antioxid. Redox Signal. 2011, 15, 2669–2681. [Google Scholar] [CrossRef] [PubMed]
  411. Choi, J.; Oh, T.G.; Jung, H.W.; Park, K.Y.; Shin, H.; Jo, T.; Kang, D.S.; Chanda, D.; Hong, S.; Kim, J.; et al. Estrogen-Related Receptor γ Maintains Pancreatic Acinar Cell Function and Identity by Regulating Cellular Metabolism. Gastroenterology 2022, 163, 239–256. [Google Scholar] [CrossRef] [PubMed]
  412. Shirihai, O.S.; Song, M.; Dorn, G.W., 2nd. How mitochondrial dynamism orchestrates mitophagy. Circ. Res. 2015, 116, 1835–1849. [Google Scholar] [CrossRef] [PubMed]
  413. Vanasco, V.; Ropolo, A.; Grasso, D.; Ojeda, D.S.; García, M.N.; Vico, T.A.; Orquera, T.; Quarleri, J.; Alvarez, S.; Vaccaro, M.I. Mitochondrial Dynamics and VMP1-Related Selective Mitophagy in Experimental Acute Pancreatitis. Front. Cell Dev. Biol. 2021, 9, 640094. [Google Scholar] [CrossRef] [PubMed]
  414. Wen, E.; Xin, G.; Su, W.; Li, S.; Zhang, Y.; Dong, Y.; Yang, X.; Wan, C.; Chen, Z.; Yu, X.; et al. Activation of TLR4 induces severe acute pancreatitis-associated spleen injury via ROS-disrupted mitophagy pathway. Mol. Immunol. 2022, 142, 63–75. [Google Scholar] [CrossRef] [PubMed]
  415. Zhi, X.; Feng, W.; Rong, Y.; Liu, R. Anatomy of autophagy: From the beginning to the end. Cell Mol. Life Sci. 2018, 75, 815–831. [Google Scholar] [CrossRef] [PubMed]
  416. Savini, M.; Zhao, Q.; Wang, M.C. Lysosomes: Signaling Hubs for Metabolic Sensing and Longevity. Trends Cell Biol. 2019, 29, 876–887. [Google Scholar] [CrossRef]
  417. Nowosad, A.; Besson, A. Lysosomes at the Crossroads of Cell Metabolism, Cell Cycle, and Stemness. Int. J. Mol. Sci. 2022, 23, 2290. [Google Scholar] [CrossRef] [PubMed]
  418. Reiser, J.; Adair, B.; Reinheckel, T. Specialized roles for cysteine cathepsins in health and disease. J. Clin. Investig. 2010, 120, 3421–3431. [Google Scholar] [CrossRef] [PubMed]
  419. Eskelinen, E.L.; Tanaka, Y.; Saftig, P. At the acidic edge: Emerging functions for lysosomal membrane proteins. Trends Cell Biol. 2003, 13, 137–145. [Google Scholar] [CrossRef] [PubMed]
  420. Saftig, P.; Klumperman, J. Lysosome biogenesis and lysosomal membrane proteins: Trafficking meets function. Nat. Rev. Mol. Cell Biol. 2009, 10, 623–635. [Google Scholar] [CrossRef]
  421. Saluja, A.; Hashimoto, S.; Saluja, M.; Powers, R.E.; Meldolesi, J.; Steer, M.L. Subcellular redistribution of lysosomal enzymes during caerulein-induced pancreatitis. Am. J. Physiol. 1987, 253 Pt 1, G508–G516. [Google Scholar] [CrossRef] [PubMed]
  422. Tan, A.; Prasad, R.; Lee, C.; Jho, E.H. Past, present, and future perspectives of transcription factor EB (TFEB): Mechanisms of regulation and association with disease. Cell Death Differ. 2022, 29, 1433–1449. [Google Scholar] [CrossRef] [PubMed]
  423. Napolitano, G.; Ballabio, A. TFEB at a glance. J. Cell Sci. 2016, 129, 2475–2481. [Google Scholar] [CrossRef] [PubMed]
  424. Wang, S.; Ni, H.M.; Chao, X.; Wang, H.; Bridges, B.; Kumer, S.; Schmitt, T.; Mareninova, O.; Gukovskaya, A.; De Lisle, R.C.; et al. Impaired TFEB-mediated lysosomal biogenesis promotes the development of pancreatitis in mice and is associated with human pancreatitis. Autophagy 2019, 15, 1954–1969. [Google Scholar] [CrossRef] [PubMed]
  425. Wang, S.; Ni, H.M.; Chao, X.; Ma, X.; Kolodecik, T.; De Lisle, R.; Ballabio, A.; Pacher, P.; Ding, W.X. Critical Role of TFEB-Mediated Lysosomal Biogenesis in Alcohol-Induced Pancreatitis in Mice and Humans. Cell Mol. Gastroenterol. Hepatol. 2020, 10, 59–81. [Google Scholar] [CrossRef] [PubMed]
  426. Mizushima, N. The ubiquitin E2 enzyme UBE2QL1 mediates lysophagy. EMBO Rep. 2019, 20, e49104. [Google Scholar] [CrossRef] [PubMed]
  427. Maejima, I.; Takahashi, A.; Omori, H.; Kimura, T.; Takabatake, Y.; Saitoh, T.; Yamamoto, A.; Hamasaki, M.; Noda, T.; Isaka, Y.; et al. Autophagy sequesters damaged lysosomes to control lysosomal biogenesis and kidney injury. EMBO J. 2013, 32, 2336–2347. [Google Scholar] [CrossRef] [PubMed]
  428. Iwama, H.; Mehanna, S.; Imasaka, M.; Hashidume, S.; Nishiura, H.; Yamamura, K.I.; Suzuki, C.; Uchiyama, Y.; Hatano, E.; Ohmuraya, M. Cathepsin B and D deficiency in the mouse pancreas induces impaired autophagy and chronic pancreatitis. Sci. Rep. 2021, 11, 6596. [Google Scholar] [CrossRef] [PubMed]
  429. Yuan, X.; Wu, J.; Guo, X.; Li, W.; Luo, C.; Li, S.; Wang, B.; Tang, L.; Sun, H. Autophagy in Acute Pancreatitis: Organelle Interaction and microRNA Regulation. Oxid. Med. Cell Longev. 2021, 2021, 8811935. [Google Scholar] [CrossRef] [PubMed]
  430. Sun, H.; Tian, J.; Li, J. MiR-92b-3p ameliorates inflammation and autophagy by targeting TRAF3 and suppressing MKK3-p38 pathway in caerulein-induced AR42J cells. Int. Immunopharmacol. 2020, 88, 106691. [Google Scholar] [CrossRef]
  431. Xiao, J.; Feng, X.; Huang, X.Y.; Huang, Z.; Huang, Y.; Li, C.; Li, G.; Nong, S.; Wu, R.; Huang, Y.; et al. Spautin-1 Ameliorates Acute Pancreatitis via Inhibiting Impaired Autophagy and Alleviating Calcium Overload. Mol. Med. 2016, 22, 643–652. [Google Scholar] [CrossRef] [PubMed]
  432. Xiao, J.; Lin, H.; Liu, B.; Jin, J. CaMKII/proteasome/cytosolic calcium/cathepsin B axis was present in tryspin activation induced by nicardipine. Biosci. Rep. 2019, 39, BSR20190516. [Google Scholar] [CrossRef] [PubMed]
  433. Zalcman, G.; Federman, N.; Romano, A. CaMKII Isoforms in Learning and Memory: Localization and Function. Front. Mol. Neurosci. 2018, 11, 445. [Google Scholar] [CrossRef] [PubMed]
  434. Ji, L.; Wang, Z.H.; Zhang, Y.H.; Zhou, Y.; Tang, D.S.; Yan, C.S.; Ma, J.M.; Fang, K.; Gao, L.; Ren, N.S.; et al. ATG7-enhanced impaired autophagy exacerbates acute pancreatitis by promoting regulated necrosis via the miR-30b-5p/CAMKII pathway. Cell Death Dis. 2022, 13, 211. [Google Scholar] [CrossRef] [PubMed]
  435. Yu, J.H.; Kim, H. Role of janus kinase/signal transducers and activators of transcription in the pathogenesis of pancreatitis and pancreatic cancer. Gut Liver 2012, 6, 417–422. [Google Scholar] [CrossRef] [PubMed]
  436. Yang, S.; Imamura, Y.; Jenkins, R.W.; Cañadas, I.; Kitajima, S.; Aref, A.; Brannon, A.; Oki, E.; Castoreno, A.; Zhu, Z.; et al. Autophagy Inhibition Dysregulates TBK1 Signaling and Promotes Pancreatic Inflammation. Cancer Immunol. Res. 2016, 4, 520–530. [Google Scholar] [CrossRef] [PubMed]
  437. Kubisch, C.H.; Logsdon, C.D. Endoplasmic reticulum stress and the pancreatic acinar cell. Expert Rev. Gastroenterol. Hepatol. 2008, 2, 249–260. [Google Scholar] [CrossRef] [PubMed]
  438. Sah, R.P.; Garg, S.K.; Dixit, A.K.; Dudeja, V.; Dawra, R.K.; Saluja, A.K. Endoplasmic reticulum stress is chronically activated in chronic pancreatitis. J. Biol. Chem. 2014, 289, 27551–27561. [Google Scholar] [CrossRef] [PubMed]
  439. Zhang, K.; Kaufman, R.J. From endoplasmic-reticulum stress to the inflammatory response. Nature 2008, 454, 455–462. [Google Scholar] [CrossRef] [PubMed]
  440. Netea-Maier, R.T.; Plantinga, T.S.; van de Veerdonk, F.L.; Smit, J.W.; Netea, M.G. Modulation of inflammation by autophagy: Consequences for human disease. Autophagy 2016, 12, 245–260. [Google Scholar] [CrossRef]
  441. Martins, J.D.; Liberal, J.; Silva, A.; Ferreira, I.; Neves, B.M.; Cruz, M.T. Autophagy and inflammasome interplay. DNA Cell Biol. 2015, 34, 274–281. [Google Scholar] [CrossRef] [PubMed]
  442. Zhang, T.; Gan, Y.; Zhu, S. Association between autophagy and acute pancreatitis. Front. Genet. 2023, 14, 998035. [Google Scholar] [CrossRef] [PubMed]
  443. Hey-Hadavi, J.; Velisetty, P.; Mhatre, S. Trends and recent developments in pharmacotherapy of acute pancreatitis. Postgrad. Med. 2023, 135, 334–344. [Google Scholar] [CrossRef] [PubMed]
  444. Zaman, S.; Gorelick, F. Acute pancreatitis: Pathogenesis and emerging therapies. J. Pancreatol. 2024, 7, 10–20. [Google Scholar] [CrossRef]
  445. Yang, H.; Ma, S.; Guo, Y.; Cui, D.; Yao, J. Bidirectional effects of pyrrolidine dithiocarbamate on severe acute pancreatitis in a rat model. Dose. Response 2019, 17, 1559325819825905. [Google Scholar] [CrossRef] [PubMed]
  446. Wan, J.; Chen, J.; Wu, D.; Yang, X.; Ouyang, Y.; Zhu, Y.; Xia, L.; Lu, N. Regulation of autophagy affects the prognosis of mice with severe acute pancreatitis. Dig. Dis. Sci. 2018, 63, 2639–2650. [Google Scholar] [CrossRef] [PubMed]
  447. Yang, S.; Bing, M.; Chen, F.; Sun, Y.; Chen, H.; Chen, W. Autophagy regulation by the nuclear factor kB signal axis in acute pancreatis. Pancreas 2012, 41, 367–373. [Google Scholar] [CrossRef] [PubMed]
  448. Fu, X.; Xiu, Z.; Xu, H. Interleukin-22 and acute pancreatitis: A review. Medicine 2023, 102, e35695. [Google Scholar] [CrossRef] [PubMed]
  449. Fu, X.; Xiu, Z.; Xu, Q.; Yue, R.; Xu, H. Interleukin-22 Alleviates Caerulein-Induced Acute Pancreatitis by Activating AKT/mTOR Pathway. Dig. Dis. Sci. 2024; Online ahead of print. [Google Scholar] [CrossRef]
  450. Dong, K.; Chen, X.; Xie, L.; Yu, L.; Shen, M.; Wang, Y.; Wu, S.; Wang, J.; Lu, J.; Wei, G.; et al. Spautin-A41 Attenuates Cerulein-Induced Acute Pancreatitis through Inhibition of Dysregulated Autophagy. Biol. Pharm. Bull. 2019, 42, 1789–1798. [Google Scholar] [CrossRef] [PubMed]
Figure 1. A simplified diagram of autophagy regulation. Black arrows: activation. Red arrows: inhibition. Intermittent arrows: cleavage. Certain pathways have been omitted for clarity. See text for more details. Bcl-2: B-cell lymphoma-2; FADD: Fas-associating protein with death domain; TRADD: Tumor necrosis factor receptor type 1-associated DEATH domain protein; RIPK1: Receptor Interacting Serine/Threonine Kinase 1.
Figure 1. A simplified diagram of autophagy regulation. Black arrows: activation. Red arrows: inhibition. Intermittent arrows: cleavage. Certain pathways have been omitted for clarity. See text for more details. Bcl-2: B-cell lymphoma-2; FADD: Fas-associating protein with death domain; TRADD: Tumor necrosis factor receptor type 1-associated DEATH domain protein; RIPK1: Receptor Interacting Serine/Threonine Kinase 1.
Gastroent 15 00022 g001
Figure 2. Pathogenesis of acute pancreatitis. Black arrows: activation. Red arrows: inhibition. ER: endoplasmic reticulum; MPTP: mitochondrial permeability transition pores; ATP: adenosine triphosphate; CHOP: CEBP homologous protein; DAMPS: damage associated molecular patterns; ETOH: Alcohol; SERCA: smooth ER Ca++ channels; PMCA: plasma membrane Ca++ channels; UPR: unfolded protein response.
Figure 2. Pathogenesis of acute pancreatitis. Black arrows: activation. Red arrows: inhibition. ER: endoplasmic reticulum; MPTP: mitochondrial permeability transition pores; ATP: adenosine triphosphate; CHOP: CEBP homologous protein; DAMPS: damage associated molecular patterns; ETOH: Alcohol; SERCA: smooth ER Ca++ channels; PMCA: plasma membrane Ca++ channels; UPR: unfolded protein response.
Gastroent 15 00022 g002
Figure 3. Inflammation in acute and chronic pancreatitis. Black arrows: activation. For more details, see text. LPS: lipopolysaccharide; HSP70: heat shock protein 70; HMGB1: high mobility group box 1; TLR4,9: toll like receptor 4,9; ROS: reactive oxygen species; ATP: adenosine triphosphate; NETS: neutrophil extracellular traps; NLRP3: NLR pyrin domain containing protein 3; ASC: caspase recruitment domain; NOD1: nucleotide-binding oligomerization domain 1; MCP1: monocyte chemoattractant protein 1; mtDNA: mitochondrial DNA; PSC: Pancreatic stellate cells; MIP2: Macrophage inflammatory protein-2.
Figure 3. Inflammation in acute and chronic pancreatitis. Black arrows: activation. For more details, see text. LPS: lipopolysaccharide; HSP70: heat shock protein 70; HMGB1: high mobility group box 1; TLR4,9: toll like receptor 4,9; ROS: reactive oxygen species; ATP: adenosine triphosphate; NETS: neutrophil extracellular traps; NLRP3: NLR pyrin domain containing protein 3; ASC: caspase recruitment domain; NOD1: nucleotide-binding oligomerization domain 1; MCP1: monocyte chemoattractant protein 1; mtDNA: mitochondrial DNA; PSC: Pancreatic stellate cells; MIP2: Macrophage inflammatory protein-2.
Gastroent 15 00022 g003
Table 1. Main studies on the role of autophagy in pancreatitis.
Table 1. Main studies on the role of autophagy in pancreatitis.
Original StudiesOutcomesReferences
Deletions of Atg5 or Atg7 or of the inhibitor of nuclear factor IκB kinase α (IKKα) ER stress and accumulation of dysfunctional mitochondria unable to generate ATP[178,271]
Atg5 deletion Reduced severity of the disease paralleled with the reduced trypsinogen activation[393,394]
LAMP2 deficiency Increased severity of cerulein pancreatitis[270,374]
Administration of the enhancer of autophagy trehalose Reduced trypsinogen activation and necrosis[122]
Reduced autophagy in severe APImpaired tight junctions. Reduction of the function of goblet and Paneth cells. Increased bacterial translocation and extra-pancreatic manifestations[370,378,379]
Zymogen exocytosis and autophagy. SNARE proteinsBlock of the fusion of zymogen granules with the plasma membrane and exocytosis[382,383]
Pancreatic Protein kinase C iota (PKCi) deletionDisruption of autophagy. Increased sensitivity to cerulein-induced pancreatitis [385]
Stimulation of autophagic flux by the FXR-OSGIN1 axisProtection from pancreatitis[389]
Increased zymophagyProtection from pancreatitis[390,391]
Rab9 decrease Boost of canonical autophagy and mitigation of disease severity [392]
Xanthohumol administrationInhibition of mTOR. Restoration of autophagy. Reduction of pancreatitis severity[384]
ER stressActivation of trypsinogen and impaired autophagy[402,403,404,407]
ROS overproductionMitophagy disruption. Activation of AKT/mTOR pathway. Severe AP[414]
[407] Dysfunction of the lysosomes Autophagy block. Pancreatitis[376,395,419]
Deletion/degradation of TFEBAutophagy impairment. Increased severity of pancreatitis.[424,425]
The associations between autophagy and pancreatitis were recently reviewed [442].
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tsomidis, I.; Voumvouraki, A.; Kouroumalis, E. The Pathogenesis of Pancreatitis and the Role of Autophagy. Gastroenterol. Insights 2024, 15, 303-341. https://doi.org/10.3390/gastroent15020022

AMA Style

Tsomidis I, Voumvouraki A, Kouroumalis E. The Pathogenesis of Pancreatitis and the Role of Autophagy. Gastroenterology Insights. 2024; 15(2):303-341. https://doi.org/10.3390/gastroent15020022

Chicago/Turabian Style

Tsomidis, Ioannis, Argyro Voumvouraki, and Elias Kouroumalis. 2024. "The Pathogenesis of Pancreatitis and the Role of Autophagy" Gastroenterology Insights 15, no. 2: 303-341. https://doi.org/10.3390/gastroent15020022

Article Metrics

Back to TopTop