Next Article in Journal
Human RPF1 and ESF1 in Pre-rRNA Processing and the Assembly of Pre-Ribosomal Particles: A Functional Study
Previous Article in Journal
Fourier Ptychographic Microscopy 10 Years on: A Review
Previous Article in Special Issue
MiR-630 Promotes Radioresistance by Induction of Anti-Apoptotic Effect via Nrf2–GPX2 Molecular Axis in Head–Neck Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cell Death, by Any Other Name…

by
Mustapha Kandouz
1,2
1
Department of Pathology, School of Medicine, Wayne State University, 540 East Canfield Avenue, Detroit, MI 48201, USA
2
Karmanos Cancer Institute, Wayne State University, Detroit, MI 48201, USA
Cells 2024, 13(4), 325; https://doi.org/10.3390/cells13040325
Submission received: 31 December 2023 / Revised: 4 February 2024 / Accepted: 6 February 2024 / Published: 10 February 2024

Abstract

:
Studies trying to understand cell death, this ultimate biological process, can be traced back to a century ago. Yet, unlike many other fashionable research interests, research on cell death is more alive than ever. New modes of cell death are discovered in specific contexts, as are new molecular pathways. But what is “cell death”, really? This question has not found a definitive answer yet. Nevertheless, part of the answer is irreversibility, whereby cells can no longer recover from stress or injury. Here, we identify the most distinctive features of different modes of cell death, focusing on the executive final stages. In addition to the final stages, these modes can differ in their triggering stimulus, thus referring to the initial stages. Within this framework, we use a few illustrative examples to examine how intercellular communication factors in the demise of cells. First, we discuss the interplay between cell–cell communication and cell death during a few steps in the early development of multicellular organisms. Next, we will discuss this interplay in a fully developed and functional tissue, the gut, which is among the most rapidly renewing tissues in the body and, therefore, makes extensive use of cell death. Furthermore, we will discuss how the balance between cell death and communication is modified during a pathological condition, i.e., colon tumorigenesis, and how it could shed light on resistance to cancer therapy. Finally, we briefly review data on the role of cell–cell communication modes in the propagation of cell death signals and how this has been considered as a potential therapeutic approach. Far from vainly trying to provide a comprehensive review, we launch an invitation to ponder over the significance of cell death diversity and how it provides multiple opportunities for the contribution of various modes of intercellular communication.

1. Introduction: The Dead Cell

It is now customary in the field of cell death to trace back its earliest mentions to more than a century ago, in fact, to the 19th century German naturalist Carl Vogt and his work on the nervous system of toad embryos or even to earlier observations [1]. Many decades later, histological studies of ischemic liver injury carried out between 1962 and 1964 were behind the distinction between classical necrotic cell death and what will later be known as apoptotic cell death or simply apoptosis [2].
Although every type of cell death proceeds through different stages, it is when it has reached the point of no return, where the decay inflicted to it is irreversible, that a cell would be considered dead. A group of leading researchers in the field, part of the “Nomenclature Committee on Cell Death”, has proposed a set of molecular and morphological criteria to identify a “dead cell” [3]. So, apart from initiatory and regulatory mechanisms, what exactly are these landmark morphological and biochemical events that define the essence of cell death? We know that cells can recover from apoptosis up to a certain point [4,5,6,7,8,9,10,11], and a name has been coined for that mechanism of recovery: anastasis, from the Greek “rising to life”, also known as “resurrection” [5]. While saving cells from certain death is a desirable outcome in many situations, particularly in degenerative diseases, it might come at a highly dangerous price, as it allows cells which have acquired mutations to undergo oncogenic transformation [12,13,14]. A profiling study suggested that, even during apoptosis, cells express a “poise for recovery” set of genes encoding survival proteins that, although not translated, can help cells to recover if the apoptotic stress is removed [15]. In fact, cells have a “life”, at least biochemically speaking, after death [16], and some processes persist even after the cell’s remnants have been engulfed by phagocytosis, as is the case with the release of lysosomal cathepsins into the cytosol [17,18]. The generated apoptotic bodies (ABs) even acquire a second identity as intercellular communication moieties, similar to exosomes and other extracellular vesicles [19,20,21]. The challenge regarding any event occurring beyond the point where anastasis is impossible is to know whether it is a consequence of death or an active executioner of it. However, a more pressing aim for us here is to have an idea about the events surrounding the border between the life and death of the cell.
The accidental death of cells known as necrosis (from the Greek etymology nekrōsis, “nekros” meaning (dead) corpse and “-osis” meaning a process or condition) appears as the swelling and rupture of the organelles and cell membrane, and, in the absence of provisions for cleaning and disposal, the process results in the “messy” spillage of intracellular contents in the extracellular environment. Necrosis can result from a large array of acute injuries and extreme stresses, such as exposure to drastic changes in temperatures or pH, radiation, toxins, osmotic shock, pressure, shear force, etc. There is a trending set of data in support of considering necrosis not as an accidental non-physiological type of cell death but as a molecularly defined and normal process [22,23]. Furthermore, as a variant on the theme of the above-mentioned primary necrosis, a so-called secondary necrosis can become, under some circumstances, the fate of cells dying by other death processes such as apoptosis [24,25] (Figure 1). Indeed, when a terminally apoptotic cell is not cleared adequately, leakage of its intracellular content can trigger an inflammatory response and secondary necrosis [26,27]. In fact, the same event can either lead to apoptosis or necrosis. For instance, lysosomal disruption results in the release of cathepsin proteases into the cytosol and cellular autodigestion. Depending on the nature of the substrate of these cathepsins, what ensues is either apoptosis or necrosis [28]. Other instances of regulated cell death also elicit inflammation. These include necroptosis and pyroptosis, induced by pathogen infection. Necroptosis is a type of regulated necrotic cell death that, unlike accidental necrotic cell death, can be prevented by caspase inhibitors [29,30,31]. In this, necroptosis resembles caspase-dependent apoptosis. Necroptotic cell death is characterized by the formation of cytoplasmic membrane pore complexes through which damage-associated molecular patterns (DAMPs) are released [32]. Terminal morphological events are like those seen in accidental necrosis and include organellar swelling, cell membrane rupture, and the complete collapse of both cytoplasm and nucleus. Because it results in the release of cellular content and inflammation [33,34,35], it is understandable why this process constitutes an important defense mechanism, allowing the self-destruction of the infected cells and preventing further pathogen propagation by attracting attention from the immune cells [36]. Also, since the process is triggered by the inhibition of the caspase machinery by viral proteins and does not necessitate executioner caspases to proceed, this caspase independence, thus, sets necroptosis apart from apoptosis, although both are regulated cell death processes. Other than necroptosis, pyroptosis is also characterized by terminal necrosis-like events that trigger inflammation. Pyroptosis is a programmed necrotic cell death mechanism, specifically dependent upon cleavage of gasdermin D (GSDMD), the designated pyroptosis executioner [37], by inflammasomes and inflammatory caspases 1 and 11/4/5 [38,39,40,41]. In this, it differs from apoptosis, which uses caspases 2, 3, and 6–10. The name was coined from the word “pyro” meaning “fire” or “fever”, the fire of inflammation [39]. Pyroptosis, like necropsis, is a defense mechanism against pathogens [42,43,44]. GSDMD is responsible for forming pores in the cytoplasmic membrane. Indeed, caspases cleave off their N-terminal domain, which opens large pores in the membrane, resulting in swelling and osmotic lysis [45,46,47,48].
Phenotypically, pyroptotic cells are swollen and their membrane is ruptured, causing the release of cellular contents; their chromatin is condensed, and mitochondria lose their membrane potential and leak material [49].
Although mentions of a programmed type of cell death had appeared earlier [1,50,51], the term apoptosis, coined from the etymology “apo” meaning “from” and “ptosis” meaning “falling off”, symbolized by leaves falling off trees, was first used in 1972, as a result of a collaborative work between Drs. Kerr, Wyllie, and Currie [52]. The essential distinction here between necrosis and apoptosis is that the latter is “self-inflicted” death. This and similar anthropomorphic expressions such as “cell suicide” dominated the field for the following decades due to the findings comforting the “programmed” nature of the underlying mechanisms. Indeed, decades after its formal recognition, apoptosis remains an active area of research [53], and it actually accounts for the large majority of publications in the area of cell death (Figure 2). Apoptosis, programmed cell death (PCD) or regulated cell death (RCD), is a genetically controlled process that shows a remarkable temporally and spatially regulated sequence of events, as initially showed, for instance, in the development of the nematode Caenorhabditis elegans [54], the fruit fly Drosophila melanogaster [55,56], and mice [57,58,59]. Morphologically, apoptosis is recognizable by the presence of nuclear condensation, membrane blebbing, cell shrinkage, fragmentation, and the formation of the so-called apoptotic bodies. Therefore, while necrotic cell death releases cellular content which elicits an inflammatory reaction, the way apoptotic cellular remnants are disposed of helps escape inflammation by packaging the debris into membranous fragments that are recognized and phagocytosed [60,61,62].
Considering today’s knowledge, the discovery of differences between cell death types should be viewed as important steps in the history of understanding the process rather than defining clear-cut categories. In fact, accumulating evidence indicates that cell death consists of a continuum of morphological and biochemical events [63,64,65]. Many more other ways for a cell to die have been described, including necroptosis, ferroptosis, pyroptosis, alkaliptosis, efferocytosis, parthanatos, autosis, entotic cell death, NETotic cell death, lysosome-dependent cell death, autophagy-dependent cell death, Mitochondrial permeability transition (MPT)-driven necrosis, immunogenic cell death, cellular senescence, mitotic catastrophe, mitotic death [63,66], and more. It would be a daunting task to overview all these types and sub-types of cell death mechanisms. However, one type, autophagy-dependent cell death, deserves special emphasis, if any, in view of the phenomenal interest it has raised in the last few decades (Figure 2). From our viewing angle, it would also be impossible to talk about autophagic cell death without referring to the so-called lysosomal cell death.
Here, again, autophagy has been known for a long time. The term “autophagy” itself, from “auto” meaning “self” and “phagos” meaning “eat” (self-eating), dates back to the 1860s, but it is the discovery of the lysosome almost a century later that gave it its current biological significance [67,68], i.e., a catabolic process of lysosome/vacuole degradation of intracellular materials to produce nutrients and energy [69,70,71]. Ultrastructural observations provide the most recognizable changes, such as the increase in the number and size of autophagic vacuoles, the engulfment of mitochondria and other organelles as well as of intracellular membranes and cytoplasmic materials to be degraded upon the fusion of the autophagosomes with the lysosomes. Therefore, autophagy has always been regarded as a metabolic adaptative mechanism that allows cells to survive cell death-inducing stimuli [72,73,74]. In the presence of autophagy, cell death can occur in different forms, prominently apoptosis, or in a completely independent manner making use of the autophagy machinery [75]. While the final executive stages can interplay with the apoptotic machinery, initial events are the most distinctive aspects of the process. Although a distinctive aspect of autophagy-dependent cell death is its impact on intracellular membranes, it is ultimately its role in lysosomal degradation and the resulting changes to the metabolic and energetic equilibrium of the cell that make a difference. While the autophagosomal function does not necessarily mean the activation of apoptosis but could simply indicate a normal metabolic process [76], when unbridled, this appetite for its own components is what ultimately causes the demise of the cell. The site of the sentencing in autophagic cell death is the lysosome [77]. These organelles contain a large variety of hydrolases, mainly cathepsins, to degrade proteins as well as phospholipases to degrade phospholipids, among others.
Hydrolases, under an acidic pH, initiate a degradative process that helps recycle the cargo content. Some stress stimuli can result in the perforation of lysosomal membranes and result in lysosomal membrane permeabilization (LMP) [17,78]. In the early stages of an injury, any limited damage can be reversed thanks to the endosomal sorting complex required for the transport (ESCRT) machinery [79,80]. However, when the integrity of the lysosome is compromised and the harm done to the membrane is beyond repair, then ensues lysophagy, a process of selective autophagy-mediated lysosomal clearance. This level of damage to the lysosomes is instrumental in cell death because it results in a wide rupture and massive leakage of content. Keeping in mind that lysosomes are the “stomachs”/“recycling stations” of the cell, any spills would result in the release of both “digestive” materials such as cathepsins and also useless and harmful components and byproducts (e.g., reactive oxygen species). The released material subsequently initiates either apoptotic or necrotic cell death [17,81,82,83,84]. For instance, ROS release, by causing lipid and protein peroxidation and by damaging the lysosomal membrane, results in LMP and lysosomal cell death [77,84,85].
In addition to—and sometimes regardless of—caspases, other cysteine proteases such as calpains and cathepsins, whether aspartic or cysteine, define the executive arm of cell death [86,87]. However, key events in RCD can be executed independently from caspases. For instance, we have seen earlier that caspase-independent apoptosis can be due to the release of lysosomal cathepsins. In addition, while caspase-activated nucleases such as DFF40/CAD (40K DNA fragmentation factor or caspase-activated deoxyribonuclease) mediate apoptotic DNA degradation [88,89], in the absence of a role for caspases, chromatin condensation and chromosomal DNA fragmentation can be mediated by either the apoptosis-inducing factor (AIF) [90] or endonuclease G [91], both nucleases, upon release from the intermembrane space of the mitochondria into the cytosol.
Death by apoptosis involves massive structural dismantlement and decay of the cell, using a “death by a thousand cuts” strategy [92], which is executed by a large number of events of caspase-mediated proteolytic degradation of proteins and nuclease-mediated cleavage of nuclear DNA [93,94,95,96,97]. A large number of proteins are subject to proteolysis (c.f. caspase substrate database of caspase substrates), leading to a variety of phenotypical manifestations [98]. For instance, proteolytic cleavage of nuclear lamins, which form a filamentous nuclear cytoskeleton [99] while also having multiple non-structural functions [100], contributes to nuclear fragmentation [101,102]. Similarly, cleavage of proteins that are important for maintaining the Golgi structure, such as the Golgi matrix protein GM130 [103] or its receptor, the Golgi-stacking protein GRASP65 [104], are key events associated with Golgi fragmentation during apoptosis.
Along with cell shrinkage, apoptosis is also characterized by drastic structural and biochemical changes to the nucleus, most prominently, chromatin condensation, DNA fragmentation, degradation of the nuclear matrix and lamina, and, ultimately, nuclear fragmentation into dense bodies protected by parts of the nuclear envelope [105,106]. A complex sequence of proteolytic events mediated by caspases results in the disruption and dismantling of connections between different nuclear components and compartments, such as the detachment of nuclear membranes from the lamina fibrillar network or from condensed chromatin [107]. It is interesting that disassembly of the nuclear envelope also occurs during mitosis. However, while this process is reversible during mitosis, it is irreversible during apoptosis [107,108,109,110].
In addition to the fragmentation of the nucleus, other organelles are subject to drastic changes, including fission of the mitochondria [111,112], fragmentation of the Golgi [113,114], endoplasmic reticulum (ER) stress [115], and, of course and most spectacularly, fragmentation of the nucleus itself [105]. Dramatic fragmentation is controlled by the fusion/fission machinery and is indicative of mitochondria dysfunction [116,117]. However, the functional significance of mitochondria fragmentation that follows mitochondrial outer membrane permeabilization (MOMP) is controversial [118,119]. While MOMP is a hallmark of apoptosis, it does not always irreversibly lead to apoptosis. Also, mitochondrial membrane remodeling occurs after MOMP independently from caspase activation [120]. Furthermore, limited MOMP has been shown to trigger limited caspase activation, insufficient to trigger cell death but resulting in DNA damage and genomic instability, which contribute to cellular transformation and tumorigenesis [14].
In addition to organelles’ degradation, drastic changes also happen to the overall shape and structure of the cells, including disorganization and dissolution of the cytoskeleton and loss of volume and retraction of the cytoskeleton from the plasma membrane, all leading to shrinkage, blebbing, and the formation of apoptotic bodies [121,122,123,124,125,126,127].
In conclusion, terminal stages of cell death appear to come down to two fundamental processes, either extensive degradation and fragmentation, as seen in apoptosis, or cell lysis and leakage of content, as seen in different types of necrotic cell death (Figure 2).

2. Cell–Cell Communication: Dying Alone or Together

The importance of “sticking together” in the response of cells to external stress and injury can be seen even in organisms endowed with much less complex social life than mammals. Dictyostelium discoideum are unicellular organisms capable of aggregating into multicellular forms. Indeed, the threat of death by starvation pushes Dictyostelium cells to adhere to each other, forming a multicellular aggregate capable of migration in search of better conditions, eventually saving the majority in the form of spores, while a minority only dies [128]. This dependence of survival upon community life is testament to the importance of intercellular communication for cell life and death [129]. An important consequence of cell death is the dismantling of cell–cell contacts. Apoptosis can be induced by the inhibition of intercellular contact [130]. Reciprocally, intercellular modes can propagate cell death signals.
Cell–cell communication can be realized by many different modes. In addition to secreted chemicals such as growth factors, cells can interact via direct contact between their cytoplasmic membranes. All these modes have been implicated, in one way or another, either in responding to or conveying cell death signals. We have previously reviewed the role of intercellular communication proteins Ephs and ephrins in cell death [129]. Gap junctions and component proteins connexins and pannexins show evident yet complex functions in cell death [131,132,133]. Even the latest to join the family, including extracellular vesicles (EVs) and tunneling nanotubes (TNTs), are part of the cell communication–cell death nexus [134,135,136,137,138,139,140,141].
In the following examples, we will illustrate the complex interplay between cell death and cell–cell communication in three different settings: (1) the establishment of the rules of interplay during some key steps of early development, (2) the interplay in a pathological condition, i.e., during tumorigenesis, and (3) the role of cell–cell communication in the propagation of a cell death message and its use in therapy.

2.1. Cell Death and Group Dynamics in Early Development

Even in the absence of external stimuli, cells autonomously undergo selective cell death as part of developmental programs. In fact, billions of cells die daily to ensure proper embryonic development or physiological system function, and, sometimes, extensive cell death can be a sign of pathology and injury [142,143]. These large-scale group deaths, when necrotic, constitute an aggression that, via release of danger-associated molecular patterns (DAMPs), trigger an inflammatory and immune response. However, when part of a normal physiological condition, apoptotic deaths in large groups of cells sculpt organs, help establish multicellular networks, and ensure normal cell turnover, repair, regeneration, aging, etc. [142,144,145,146,147,148,149]. All of this happens to select cells within a community of cells in a precise and programmed manner. The question of how the various modes of cell–cell communication respond to or interfere with this selective cell death program is still widely open.
But let us go back a little in developmental time. Up until the early stages of mammalian preimplantation development, only limited structured cell–cell contacts are formed between the embryo’s cells. This is followed soon after by the establishment of different types of junctions, mainly gap junctions (GJs), adherens junctions (AJs), desmosomes, and tight junctions (TJs), which ultimately lead to the tightening of cell–cell contacts, increase in molecular exchange capabilities, and, hence, metabolic coupling [150,151,152]. The resulting metabolic homogeneity soon clears the way for heterogeneity, following ill-understood mechanisms [153,154]. In the embryo, there is an inverse relationship between metabolism and viability. According to the so-called “quiet embryo hypothesis”, embryos with a less active metabolism have a higher capacity to survive and develop because that relative inactivity generates less lethal damages [155,156,157]. In that metabolically quiet environment, under external stresses, many normal processes such as DNA damage and repair have the potential to trigger death. The body-to-be witnesses its first cell death event when it has only few cells [158,159,160]. As cells divide, the death process is accentuated in the far more multicellular and more differentiated blastocyst mainly through a wave of apoptosis in the inner cell mass (ICM), the part of the blastocyst which gives the fetus [161]. In fact, data suggest that cell death is correlated with cell number and proliferation [161]. It is worth noting that, before the establishment of membrane junctions, the first communication mode to regulate cell death is the paracrine effects of survival growth factors [160]. Also, it seems that the first cell death mode to enter in action is apoptosis. At that stage, when blastocyst cells are sensitive to apoptosis, the fate of cells that have developed gap junction intercellular communication (GJIC) becomes associated with that of their neighbors. It is suggested that GJIC is not necessary for early embryo development and that apoptosis eliminates unwanted and damaged cells before the appearance of the homogenizing action of GJIC [153].
In a metabolism-driven context, it is only normal that autophagy appears in the picture. In fact, it does so in parallel to apoptosis, as early as during gametogenesis, fertilization, and embryo preimplantation [162]. Oocytes in prepubertal rats were shown to be eliminated by the process-sharing features of apoptosis and autophagy while lacking the formation of blebs and apoptotic bodies [163]. During and shortly after egg fertilization, autophagy is responsible for the degradation of maternally inherited proteins in oocytes and for eliminating paternal mitochondrial DNA via mitophagy, thus paving the way for the synthesis of new zygotic proteins and as a prerequisite to preimplantation development [164,165,166,167,168,169]. Throughout early embryogenesis, autophagy occurs in waves where it is activated on and off. The reasons for this are not clear enough, but data suggest a role in assisting cells to dispose of the phagocytosed remnants of apoptotic cells [170,171]. The association between autophagy and cell death is further demonstrated by the fact that embryos deficient in many autophagy genes show massive cell death [172,173]. In mice at three months of age deficient for the autophagy gene Atg5 (autophagy-related 5), specifically in neural cells, granular cells undergo apoptosis. This is due to the loss of synaptic connectivity with Purkinje cells whose axons have swollen [174]. Similar results were found in mice where the autophagic FIP200 gene has been deleted [175].
Later, during early development, migratory neural crest cells are stem cells that contribute to the embryonic development of many tissues and organs in vertebrates [176,177]. Since neural crest cells are organized in sheets or streams which move in multicellular groups [178], it is essential for them to maintain continuous local and long-range communication with each other [179]. This occurs via various modes, including GJIC [180,181,182,183], EVs [184], AJs [185,186], TJs [187,188,189], and Eph/ephrins [190]. However, so far, available data regarding the role of these communication modes in neural crest apoptosis only concern GJs. Gap junctions are essential for the functional coupling and migration of neural crest cells [182]. Early-migrating rat neural crest cells form functional GJ whose pharmacological inhibition results in these cells undergoing apoptosis [191]. In view of the importance of metabolic regulation in neural crests [192], it is expected that autophagy plays an essential role. This is supported, for instance, by the finding that autophagy regulates the production of neural crest cells in early chick embryos [193]. Also, autophagy activated by high glucose levels leads to neural crest apoptosis [194].

2.2. Cell Death versus Cell Communication

2.2.1. Extruding the Dead

The intestinal epithelium is among the most rapidly renewing tissues in the body. Stem cells are responsible for replenishing the epithelium in cells, as their division generates precursor cells which proliferate and differentiate, while moving away from the crypt base towards the villus tip, from where they are later shed into the lumen [195,196]. This migratory process, therefore, determines the intestinal epithelial cells’ (IECs) lifespan, and any abnormalities could have serious implications, whether in chronic inflammation-associated diseases or in tumorigenesis [197]. It has been suggested that, in the confinement of the villus tip, the high cellular density resulting from continuous proliferation induces cell death and shedding [198]. This homeostatic process is complex and involves multiple players [199,200]. Dying cells communicate with their neighbors, after which the latter undergo actin/myosin contraction to extrude out the dying cell [201]. The epithelium’s barrier function, measured by electrical resistance, is maintained during the extrusion process and so are the plasma membrane integrity, AJs, and TJs [198]. These junctions are indeed essential for maintaining the intestinal epithelial barrier during IEC shedding [202].
TJs and AJs are adhesive molecular complexes involved in intercellular communication and in the maintenance of normal cellular integrity and cell–cell barriers. TJs are essential for cell polarity in epithelial cells and paracellular permeability in endothelial tissues [203]. Loss of TJ is incriminated in many aspects of tumor progression, including polarity, differentiation, adhesion, migration, invasion, and metastasis [204,205,206,207]. TJs bind to the cytoskeleton and have an intracellular signaling function [203], mediated by an interaction with cytoplasmic adaptor proteins (e.g., zonula occludens, i.e., ZO) or transmembrane linker proteins (e.g., occludin, claudins, and junctional adhesion molecules, i.e., JAMs) [206,208,209,210,211,212,213]. AJs are also important for epithelial tissues and use cadherins and related proteins to connect cytoskeletal structures between cells. Through their cytoplasmic domains, cadherins associate with actin filaments-based cytoskeletal structures [214,215]. AJs link the actin filaments of interacting cells at sites of cell–cell adhesion [216]. TJ rearrangement is important for the process of extrusion of dying cells from the epithelium and barrier maintenance [217]. Although not formally addressed, the fact that IECs’ death involves both a cytoskeletal reorganization and the maintenance of junctional integrity and function directly implicates these junctions in the process of cell death sensing and extrusion of the dying cell [198,201]. Interestingly, the latter communicates out its signal to request extrusion early in the apoptotic process, even before caspase activation and the appearance of apoptotic morphological changes or phagocytosis [201]. If anything, TJs and AJs could ensure the closing off of the intercellular gap left by the extruded cell and maintain confluence within the epithelial barrier [126].
In few words, dying cells take care of their own funerals. In the case of gut homeostasis, this is elegantly accomplished via soliciting the cytoskeletal connection with cell–cell junctions, to extrude them, while leaving behind a preserved epithelial barrier integrity.

2.2.2. Setting Fire

One cannot talk about intestinal epithelium homeostasis without talking about inflammation. In fact, the gastrointestinal tract is a place where cell death, intercellular communication, and inflammation are intimately linked.
The biggest threat to the integrity of the modes of cell–cell communication does not come from within the epithelium. Enteric bacterial pathogens have tremendous effects on junction barriers tasked with keeping them at bay. They try to alter these junctions in order to increase epithelial permeability and, whenever possible, to cross the barrier altogether [218]. In turn, epithelial cell turnover can be increased as a mechanism of expelling pathogens [219]. Such interactions between pathogens and the intestinal epithelium are under the control of an inflammatory process that, when interrupted, could lead to higher levels of apoptosis of IECs, the impaired production of antimicrobial peptides, and, ultimately, the invasion of the mucosa by pathogens [220].
Inflammatory cytokines released by immune cells constitute a threat to the intestinal barrier’s integrity, particularly in conditions such as those found in inflammatory bowel disease [221]. However, this effect of inflammation on increased permeability, in many instances, does not involve apoptosis [222]. In fact, as we have seen above, apoptotic cells signal to their neighbors to extrude them and rearrange TJs to close the gaps [201,202].
Cell death of IECs has been reported to occur not only via apoptosis but also through inflammation-regulated necrosis, necroptosis, and pyroptosis [223]. Excessive apoptosis has been associated with inflammation in the intestine [224]. While this can result from a pathological condition, inflammation in IECs is also part of a mechanism of host protection against potentially harmful enteric pathogens. Another mechanism through which apoptosis could promote inflammation and immunity is by releasing EVs [225,226]. Many mechanisms are at play. For instance, EVs can participate in antigen presentation [226,227]. In addition, extruded apoptotic cells release ABs that, upon engulfment, stimulate cytokine production by the phagocyting cell [228]. Necroptosis has also been involved in intestinal inflammation and tumorigenesis [229,230]. Pyroptosis is also an important part of the host defense arsenal in epithelial tissues [231]. Activation of the inflammasome has been involved in IEC extrusion, along with a loss of plasma membrane integrity and death by pyroptosis. Here, again, the expulsion process is associated with major actin rearrangement in neighboring cells, tasked with maintaining the intestinal epithelium’s integrity [232]. Gasdermins (GSDMs), the principal effectors of pyroptosis [39,41,63,233] which also regulate mitochondrial oxidative stress [234] and are involved in the regulation of autophagy [235,236], have been assigned multiple functions in gut inflammation [237]. While autophagy is important for intestinal epithelium homeostasis [238] and, in fact, for the host’s defense against pathogens [239], it counteracts inflammation [240,241], a function which is important in inflammatory pathogenesis such as in inflammatory bowel diseases [242].
In conclusion, the regulation of the epithelial barrier’s integrity, performed by TJs and AJs, involves a mechanism of extrusion of the dying cells. Pathogens constitute a threat to this process, and it is only normal that, in the response of the host gut to this challenge, inflammatory processes dominate. Therefore, the recourse of the intestinal epithelium to pyroptosis and other inflammation-dependent forms of cell death makes the most sense, even if this entails a risk of perturbing cell–cell interactions and damaging the epithelial barrier’s integrity.

2.2.3. Trying to Fit In

In the intestine, two systems of intercellular communication work in concert to accompany the course of cell renewal, migration, and differentiation. While TJs and AJs are important for the integrity of the epithelial layer, the Wnt/β-catenin/Tcf pathway regulates cell positioning in the crypt in addition to coordinating intestinal stem cells’ migration and proliferation, via Ephs and ephrins [243,244]. And, while the junctional mode has a critical impact on inflammatory diseases, research has proven a tremendous impact of Ephs/ephrins on tumorigenesis.
Ephs/ephrins constitute, in fact, the largest subfamily of receptor tyrosine kinases (RTKs), a fact which fueled interest in using them as targets in cancer therapy [245,246,247]. Ephs and ephrins were initially described as guidance molecules, although it is their discovery in cancer cells which gave them their name (Erythropoietin-producing human hepatocellular carcinoma). This is a large family that enlists both receptors and their ligands called ephrins. In mammals, there are fourteen Eph receptors interacting with eight Ephrin ligands. Both Eph receptors and ephrin ligands are membrane-bound, thus eliciting a bidirectional signaling in the interacting pair of cells [248]. They constitute an intercellular communication mode with many functions, such as the formation of spatial boundaries during normal development [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20]. While they regulate processes such as proliferation, cell death, and invasion, it is their role in cell sorting and positioning that made their fame [249,250], for instance, in regulating the positioning of intestinal epithelial cells within the stem cell niche [244] and coordinating between intercellular communication, migration, and cell positioning [243].
EphB2 and EphB3, two members of the Eph/ephrin family, are major elements of the genetic module that controls the compartmentalization of epithelial cells along the crypt axis and are involved in the regulation of their ordered migration [243].
In recent years, a role for Ephs and Ephrins deregulations has emerged in cancer progression and metastasis [251,252,253]. The EphB2 receptor has been reported as a tumor suppressor in the colon [254]. The initial observation was that EphBs are transcriptionally regulated by the β-catenin/Tcf signaling pathway, a major player in human colorectal cancer progression [255,256]. While cells in early lesions of dysplastic crypts and small adenomas, like normal crypt progenitor cells, were found to express EphB2, high-grade tumor areas contained EphB2-negative cells [254,255,256]. Therefore, colon cancer progression is associated with a loss of EphB2 expression. This was further confirmed in the ApcMin/+ mouse model of hereditary colon cancer and in many sporadic cancers [257]. Colorectal neoplasms in the ApcMin/+ mice usually fail to cross the adenoma-carcinoma transition. Crossing ApcMin/+ mice with animals expressing a dominant negative form of EphB2 resulted in a loss of EphB activity that accelerated colorectal tumor progression. In addition, high levels of EphB2 expression were found to be associated with a longer mean duration of survival in colorectal cancer [258], and a loss of EphB2 expression has been reported in colorectal tumors [259,260]. These findings support a tumor suppressor function of EphB2.
Tumor suppression can be achieved by inhibiting cell proliferation or inducing cell death or differentiation. Many Eph/ephrin family members have functions in cell death [129]. EphB2 has been found to regulate the proliferation of intestinal stem cells [244]. Furthermore, EphBs’ role in tumor suppression is performed by limiting the expansion of colorectal cancer cells to specific compartments, rendering difficult the incursion of EphB-expressing cells into areas of high repulsive forces from normal ephrinB-expressing intestinal cells [255]. In fact, a decreasing gradient of EphB2 expression, from the bottom to the top, is found in the proliferative crypt of the small intestine. In the large intestine, EphB2 is expressed only in the progenitor cells at the crypt base. By contrast, the EphB ligands EphrinsB1 and B2 are expressed, also as a gradient, by the differentiated surface and villus cells [255]. During tumorigenesis, the loss of EphB receptors relieves cells from spatial restriction and allows cells to intermingle freely and invade [261]. The tumor suppression effect of EphBs, by controlling the compartmentalization of tumor cells, depends on E-cadherin-mediated adhesion [261]. EphB/ephrinB interaction has also been shown to promote mesenchymal–epithelial transition (MET), reorganizing the cytoskeleton and restoring epithelial E-cadherin/ZO-1-based cell–cell junctions [262]. These effects were also accompanied by the induction of apoptosis [262].
Autophagy is important in intestinal biology through controlling the TJs barrier function, as a survival mechanism, and by regulating intestinal stem cells’ metabolism [263]. Although the data on the function of Ephs/ephrins in autophagy are scarce, the available information supports a direct connection. We have previously shown that EphB2 regulates an autophagy-dependent cell death [264,265]. EphB/ephrinB interaction was shown to use autophagy to clear IECs of intracellular pathogens [266]. Silencing of EphA1 and EphB2 was shown to block autophagy and cell death in colorectal cancer cells [267]. Further investigation is needed.
And now, a few words about the function of Ephs and ephrins in inflammation and immunity that has increasingly been acknowledged [268,269]. The role of these proteins in angiogenesis is one of the first and most studied aspects of their biological functions [270]. A major part of the function of Ephs/ephrins in gut inflammation involves synaptic plasticity and the neuroimmune regulation of intestinal inflammation-related pathogenesis [268]. The relevance of the functions of Ephs/ephrins in inflammation and immunity in colorectal cancer remains, however, to be further examined. A possibility is via a role in the tumor’s immune microenvironment [246,271]. However, how cell death could be involved is an important question deserving of attention in view of the above-discussed association between inflammation and cell death.

2.2.4. Surviving Attacks

Since intercellular communication is functionally connected to cell death, it is no surprise that it plays a role in cancer therapeutic resistance. However, the extent to which the balance between the two processes, i.e., communication and death, is important for drug resistance, could be easily identified as an area of unmet need in cancer research.
Data regarding the role of Ephs and ephrins in colon cancer drug resistance is still sporadic. EphB2 was found as one of the genes whose expression is increased in colon cancer cells resistant to platinum or taxane [272]. Acquisition of resistance to cetuximab, an antibody-based EGFR inhibitor, is associated with an increase in EphB3 expression levels and EphB3/EGFR binding in colorectal cancer cells. The mechanism of resistance can be overcome by inhibiting EphB3 expression, which allows cetuximab-induced apoptosis [273]. Claudin-1 (CLDN1) promotes colorectal cancer (CRC) cells’ chemoresistance by interacting with and stabilizing EphA2. This results in enhancing the antiapoptotic AKT signaling pathway and promoting cancer stemness [274]. Directly targeting EphA2 with a tyrosine kinase inhibitor decreases cell proliferation and induces cell death in CRC cells [275,276,277]. Based on the finding that, in response to DNA damage, the ephrinB2-encoding gene is a transcriptional target of p53, a key apoptosis regulatory protein, knock down of ephrinB2 expression was used to restore apoptosis and 5-FU chemosensitivity in mutant p53-harboring CRC tumors [278]. The physical and functional association between EphA2 and EGFR plays a role in the resistance to EGFR-targeting therapies [279]. High EphA2 expression levels are associated with a worse outcome in patients treated with cetuximab [280]. Metastatic CRC cells that have NRAS (neuroblastoma RAS viral oncogene homolog)-activating mutations are resistant to cetuximab. A failure of cetuximab to downregulate EphA2 expression in NRAS-mutant CRC cells, in comparison with cetuximab-responsive CRC cells, was suggested as a potential contributor to resistance to this drug [281]. The use of the ephrinA1 ligand to activate EphA2 succeeds in restoring cetuximab activity in NRAS-mutant colorectal cells [281]. Although it is not clear whether the mechanism involves an effect on apoptosis or another type of cell death, it was shown to involve the suppression of mitogen-activated protein kinase (MAPK) and AKT hyperactivation. The roles of these pathways in both apoptosis and Ephs/ephrins signaling are established [282]. An EphA2 small-molecule inhibitor could be used successfully to reverse cetuximab resistance in CRC cells by inducing apoptosis and cell cycle G1–G2 arrest and inhibiting the MAPK and AKT pathways [283]. Although shown in a different type of cancers, i.e., head and neck squamous cell carcinoma, resistance to cetuximab and radiotherapy (RT) was associated with elevated EphB4 and ephrinB2 expression levels [284]. The therapeutic response to cetuximab-RT could be improved by concomitantly blocking the EphB4–ephrinB2 interaction that results in the inactivation of the AKT and MAPK pathways, a decrease in proliferation, and an increase in apoptosis [284]. Similarly, in non-small-cell lung cancers, silencing of ephrinB3 sensitizes cells to a combined treatment with the kinase inhibitor PKC 412 and ionizing radiation [285]. Here, again, the MAPKs and Akt signaling pathway are involved, which elicit a decreased proliferation, increased apoptosis, mitotic catastrophe, and senescence signaling [285]. These data point to a prominent role for the MAPK and AKT pathways and apoptosis. However, the link with Eph/ephrin-mediated intercellular communication awaits further direct investigation. Interestingly, MAPK inhibition along with tyrosine phosphorylation of cell junction proteins such as CLDN1 was observed in cetuximab-responsive tumors compared to their resistant counterparts [286].
While apoptosis’ prominent role in mechanisms of CRC drug resistance is well documented [287,288,289], data are in favor of an equally critical function for autophagy and autophagy-dependent cell death [290,291]. A balance between apoptosis and autophagy, controlled by p38MAPK, has an important role in resistance to 5-FU [292]. Autophagy is involved in the resistance of CRC cells to EGFR targeting using a monoclonal antibody [293]. In this case, autophagy acts as an adaptive survival mechanism because, when inhibited, the response to the drug by increasing cell death is improved using an autophagy inhibitor [293]. In fact, a combined targeting of both EGFR and autophagy is a tempting yet still underexplored strategy to bypass drug resistance in CRC [294]. To date, only a handful of publications have addressed the regulation of autophagy by Ephs and ephrins in cancer, including in CRC cells [264,265,266,275,295]. Whether this function could impact therapeutic drug responsiveness is unknown. There are even less data regarding other modes of cell death, especially ferroptosis or pyroptosis. Ferroptosis is increasingly viewed as a possible therapeutic target in CRC [296,297]. When other modes of cell death such as apoptosis fail, the induction of ferroptosis succeeds in eliminating CRC cells [298]. This is understandable, knowing that inhibition of ferroptosis by iron sequestration results in CRC progression and resistance to 5-FU [299].
In addition to Ephs and ephrins, the role of connexins in drug resistance has also been recognized, although it is sometimes paradoxical depending on whether they act via GJIC or not. Expression of some connexins has been associated with chemoresistance in a GJIC-independent manner [300,301,302]. However, GJIC can also directly contribute to these mechanisms, as in the case of fibroblasts which protect lung tumor cells from death by connecting with them via GJICs, thus resulting in cancer cells’ chemoresistance [303]. The BE mediated by GJIC increases cisplatin cytotoxicity, and the inhibition of Cx43 expression results in drug resistance [304]. In fact, the induction of connexin expression and/or the activation of GJIC have long been viewed as strategies to overcome chemoresistance or, as will be seen below, to improve the response to gene or radiation therapies. In CRC, 5-Fluorouracil (5-FU) is one of the most important chemotherapeutic drugs in use. However, resistance emerges to this drug [305]. A correlation exists between a loss of CX43 expression, metastasis, and poor prognosis, while restoring Cx43 expression leads to the suppression of CRC progression and an increase in the sensitivity to 5-fluorouracil (5-FU) [306]. In addition to 5-FU, the resistance of CRC cells to oxaliplatin and irinotecan is also associated with a loss of Cx43 expression [307]. The use of resveratrol, a natural polyphenol, has been explored to chemosensitize CRC cells to 5-FU. The mechanism involves the upregulation of cell–cell communication molecules that constitute desmosomes, gap and tight junctions, and adhesion molecules, while also enhancing the onset of apoptosis [308]. Similarly, one of the mechanisms through which all-trans retinoic acid (ATRA) is able to restore the resistance of CRC cells to paclitaxel is by improving Cx43-dependent GJIC [309]. However, this effect relies more on a role for Cx43 in reducing CRC cells’ migration and invasiveness rather than inducing apoptosis [309]. The connection between these functions of connexins and GJIC in drug resistance with various types of cell death has not been addressed specifically. Some data from other types of cancer could provide some insight. For instance, while Cx43 was reported to have a role in resistance of glioblastoma (GBM) cells to temozolomide (TMZ) thanks to an anti-apoptotic effect [302], when combined with a Cx43 inhibitor, responsiveness to TMZ is improved by inhibiting AKT signaling and inducing both autophagy and apoptosis [310]. Furthermore, under regulation by EGFR/MAPK signaling, Cx43 expression is increased in TMZ-resistant GBM cells [311]. Knocking down Cx43 expression sensitizes GBM cell to TMZ [311]. Perhaps, more spectacularly, through establishing GJIC with neighboring astrocytes, GBM cells can resist TMZ-induced apoptosis, an effect which is prevented by the knockdown of astrocytic Cx43 expression [312]. These and other data put forward the potential use of connexin and GJIC targeting as an approach to counteract TMZ resistance [313]. Connexins and autophagy have a reciprocal relation, each being both target and regulator of the other [314]. This is the case in cervical cancer cells where the overexpression of connexin32 (Cx32) promotes autophagy in a GJ-independent manner, subsequently inducing apoptosis [315]. Nevertheless, within this equation, Cx32 has another GJIC-independent function, as it suppressed apoptosis induced by combined treatment with an autophagy inhibitor and cisplatin [316]. This anti-apoptotic effect is mediated by the Cx32-driven upregulation of EGFR expression and could be abrogated using an EGFR inhibitor [316].
In summary, intercellular communication, whether via Eph/ephrin, EVs, GJs, TJs, or AJs, has essential functions in the homeostatic regeneration of the gut, its response to pathogens, its blockade of tumor progression, and its response to therapeutic agents. While some of these roles directly implicate apoptosis and, to a lesser extent, autophagy, some other roles rely more on a role as positioning, sorting, and guidance systems (Ephs and ephrins).

2.3. The “Bystander Effect”

The propagation of cell death signals has long been associated with gap junction intercellular communication (GJIC), a junction-based mode of interaction. GJs are made of intercellular channels, composed of the connexin proteins [317] which allow direct cytoplasmic continuity between cells. This structure allows the passive exchange of a plethora of small molecules, including calcium, ions, nucleotides, cAMP/cGMP, inositol 1,4,5-triphosphate, ATP, ADP, amino acids, and other metabolites and signaling molecules [318,319,320] as well as cellular electrical coupling and synchronization [321,322]. In addition to connexin-based channels, there are hemichannels based on the pannexin proteins that connect between the cytoplasm and the extracellular milieu. Both connexins and pannexins have an important role in cell death [133].
GJIC shows both pro-apoptotic [323,324,325,326,327] and anti-apoptotic [328,329,330,331,332] functions. A well-known consequence of the pro-apoptotic role of GJIC is the so-called “bystander effect” (BE), whereby a cell can transmit a cytotoxic message to its neighbors when it is connected to them via GJs [333,334]. The classical example is the use of the herpes simplex virus-thymidine kinase (HSV-tk)/ganciclovir (GCV) system as a cancer gene therapy approach. The HSV-tk gene product converts the antiviral drug GCV into GCV triphosphate, an analogue with pro-apoptotic effects. Even if only a single cell is targeted with this system, pro-apoptotic signals can be transmitted by the targeted cell, via GJs, and induce cell death in its by-standing neighbors, even if the latter do not express the necessary HSV- gene [335,336].
A similar BE is observed in cells that are exposed to ionizing radiation (IR), a long-known phenomenon with an important impact on human health [337,338,339]. Here, again, the damage inflicted upon a cell population is shared with non-irradiated neighboring cells with which they are in a GJIC [340].
In both examples of BE, the first question is about the nature of the message shared by donor cells with acceptor cells. In the gene therapy case, following the classical pro-drug activation process, GCV triphosphate analogue production is performed in the donor cell and reaches the acceptor cells. GCV triphosphate is a DNA polymerase inhibitor that is incorporated within the newly synthesized DNA, resulting in chain termination and single-strand breaks [341], thus triggering apoptosis. The model for the pro-apoptotic mechanism involves putatively a p53-dependent transcription of the CD95 receptor, which becomes activated in the absence of ligand stimulation, subsequently leading to the recruitment of the extrinsic apoptosis machinery components FADD and caspase-8, triggering the caspase cascade [342]. Similarly, radiation-associated BE involves the induction of different forms of chromosomal and genomic DNA damage, which trigger a “DNA damage response” (DDR), shared by the targeted cell with by-standing neighbors not directly subject to irradiation [343,344,345,346,347]. Once transduced, the DDR signal activates an array of molecular pathways geared towards repairing the damage [348] but eventually activating p53, p63, and p73 and eliciting apoptosis [349,350,351,352]. In addition to nuclear DNA damage, other components from the cytoplasm can be targeted by radiation [353,354], thus resulting in the generation of deleterious molecules which can also be shared by BE. These include DNA synthesis-inhibiting molecules [37] and free radicals generated by stimulated mitochondria, which can cause DNA damage [347,353]. The radiation-induced production of reactive oxygen species (ROS), an important event [355], can increase lysosomal membrane permeability, the release of acid sphingomyelinase, the activation of ceramide synthase, and, ultimately, lead to apoptosis [356,357,358,359,360].
BE-associated genotoxicity also triggers non-apoptotic forms of cell death, depending on the nature of the damage [361,362]. In addition, autophagy has been associated with a response to IR and, interfering with it, results in cell sensitization to radiation therapy [363,364,365,366]. Furthermore, irradiation could induce autophagy in both targeted and bystander cells [367] via the transfer of ROS molecules [368]. Autophagy might function as a survival response, ridding the injured cell of damaged cellular components including mitochondria, via mitophagy, a mitochondria-specific form of autophagic degradation. However, multiple data feed the ambiguity concerning the role of autophagy in the response to radiation and the BE. For instance, while autophagy protects cells from genomic instability and DNA damage [369,370], mitophagy increases radiation-induced DNA damage and cell death [371]. In addition to inducing autophagy, IR also induces ferroptosis, a type of iron-dependent and lipid peroxidation-mediated cell death mechanism [372,373]. In fact, ferroptosis is closely associated with autophagy and is, therefore, important for the metabolic machinery of cell death and survival [374,375,376,377]. Ferroptosis, however, relies heavily on lipid metabolism [377,378,379,380]. The effect of IR on ferroptosis is due to its impact on the generation of lipid peroxides and increase in iron metabolism [381]. In return, the induction of ferroptosis has been shown to sensitize cells to radiation [382]. Although different in many ways from apoptosis (absence of chromatin condensation, nuclear fragmentation, and formation of blebs) and necroptosis (no membrane rupturing) [383], ferroptosis is considered a regulated necrotic cell death. Its main morphological feature is the alteration of the mitochondria [373], and the latter’s functions are essential for this process to occur [376,384,385,386]. Although studies formally scrutinizing the link between ferroptosis and BE are lacking, there are many indications in support of this link, the first of which is the importance of mitochondria for both processes. Inhibition of the GJIC protein Cx43 blocks ferroptosis and its associated cell death [387,388]. Interestingly, depletion of the GJ protein pannexin1 attenuates lipid peroxidation and iron accumulation and, subsequently, inhibits ferroptotic cell death [389]. This is not surprising, as pannexins are involved in forming large pores and GJ hemichannels [133,390,391] and are important for signal propagation, particularly electrical coupling [392]. Furthermore, many studies have shown that ferroptosis and the cell death it generates are propagated to neighboring cells [383,393,394,395]. Therefore, BE is not restricted to apoptosis, and, in contexts where ferroptosis is favored over other cell death types, cell death propagation between targeted and non-targeted cell is possible.
In another respect, cells can share mitochondria as part of the BE. These organelles are vital for the cell’s metabolic and energetic functions and, hence, for the cell’s death and survival mechanisms. Intact mitochondria as well as fragments are known to be shared between cells via multiple communication modes, including GJs, TNTs, and EVs [396,397,398,399,400,401,402,403,404]. In doing so, healthy mitochondria can, through a fusion mechanism, restore the function of damaged mitochondria in communicating cells, thus rescuing the cell from apoptosis [136,405]. Metabolic boost by mitochondria transfer between astrocytes and glioblastoma cells increases mitochondrial respiration and the activity of metabolic pathways, thus promoting cell proliferation and tumorigenicity [406]. Cytosolic irradiation results in mitochondrial depolarization [407] and extensive fragmentation [408], launching a cascade of ROS and apoptosis. Since intracellular ROS molecules are mainly generated in the mitochondria, increasing mitochondrial fusion activity can protect cells from irradiation-induced apoptosis [408]. Using a cell co-culture system, it has been shown that healthy mitochondria transfer restores the DNA damage repair response of by-standing irradiated cells [405]. Therefore, it seems that, while irradiated cells use the BE to share cytotoxic signals with neighboring cells, the latter use the BE to send back healthy mitochondria in an attempt to restore damaged organelles and rescue the irradiation-targeted cells from death.
Although initially associated with GJIC, the BE has later been found to involve other modes of cell–cell communication as well. Grouped under the generic name of extracellular vesicles (EVs) is a large and diverse population of membrane-based vesicles released by cells into the extracellular environment [409,410,411,412,413,414,415,416,417,418]. EVs are active members of the intercellular communication toolkit and are involved in the transfer of a large variety of molecules and organelles [419,420,421,422,423,424,425,426,427,428,429,430,431,432]. Most prominent among the EVs are exosomes and apoptotic vesicles. In fact, EVs are released by many cell types in normal, stressed, or pathological conditions [83,433]. Cells exposed to IR respond by releasing exosomes [434], and it has been reported that these EVs are part of radiation-induced BE [435,436,437,438,439,440,441]. EVs appear to be particularly associated with BE signal perpetuation, as BE acceptor cells are also capable of inducing BE in other cells via EVs [437] and also by transferring microRNAs responsible for downregulating the expression of TGFβ1, which triggers BE by increasing intracellular ROS [442,443,444].
As discussed earlier, dying cells release apoptotic bodies and small vesicles, allowing them to communicate with neighboring cells [21,52,83,411,414,445,446,447,448,449,450,451,452,453]. It has been suggested that HSV-tk-targeted dying cells release apoptotic vesicles that, once taken up by nearby non-targeted acceptor cells, also enter apoptosis [454]. EVs are also capable of mitochondria transfer [402,403,404].
Irradiated cells rush to build a denser-than-normal network of TNTs [455]. TNTs can mediate the transfer of mitochondria [396,397,398,399,400,401] and are thus capable of reversing radiation-induced apoptosis at its early stages [136].
In conclusion, one of the fundamental contributions of different intercellular communication modes is the amplification of cellular signals. While this has advantages at the tissular level to ensure a level of intercellular coordination and synchronization, it can become detrimental to the cells when it is taken advantage of, as exemplified by the BE cytotoxic effects. The BE has indeed important biological and clinical implications, particularly in disease therapy [346,456,457].

3. Concluding Remarks

The question of what constitutes a dead cell has been and remains an open and complex one. For decades after the discovery of necrosis and apoptosis, pharmacological inhibition and genetic loss of function experiments have reached (the intriguingly obvious) conclusion that irreversibility is the defining parameter. Therefore, damages to the mitochondria and possibly some of the executioner caspase-driven proteolytic events would be essential indicators. However, it is now clear that what defines irreversibility is far from being ubiquitous. Cells can survive caspase activation or mitochondrial loss of integrity following a process called anastasis [8]. Interestingly, it has been shown that early recovery from apoptotic stress was associated with entry in a proliferative state followed by migration [15]. These data suggest that the irreversibility of cell death, at least via apoptosis, depends on the ability of cells to engage in a different homeostatic process, e.g., from growth arrest to proliferation and from the latter to migration. Indeed, multicellular organisms have developed a coupling between homeostatic processes that is lacking in less developed organisms such as Dictyostelium discoideum [128]. Further investigation of this coupling hypothesis would have an important impact particularly for the understanding of pathological processes whereby getting rid of diseased cells is an objective that can be challenged by recovery from and resistance to cell death.
Although many modes of cell death processes have been described over the years, it appears that they share many executive mechanisms and differ mainly in their morphological features, their triggers, and whether they are strong inducers of an inflammatory response. Another distinction is the way in which specific modes interplay with other homeostatic processes such as proliferation, migration, and differentiation. To these distinctions, we would like to suggest adding how each cell death process interplays with cell–cell communication modes. The role of intercellular communication in propagating cell death signals has been a very promising axis of research, particularly towards developing gene and radiation therapies. While this has mainly focused on the GJIC-mediated bystander effect, understanding the role of other modes of communication could provide a better refinement of therapeutic strategies.
Cell–cell communication is critical for the survival and growth of multicellular organisms. In addition to growth factors and other chemical moieties, different modes of membrane-based interactions exist that are endowed with important functions and that, when defective or lost, can have dramatic impacts on cell life or death and prove instrumental in human diseases such as cancer [129,458]. It goes without saying that, considering that these interactions happen at the gate of cells, we are more concerned here with the initiation phase of cell death rather than its later phases. While there is a variety of modes of cell death, there is also a large variety of cell–cell communication modes. Cell death and intercellular communication are intertwined along the normal developmental stages of organisms and remain so in pathological conditions. On the brink of death or even after death, as a last message to the cellular world, a cell dying by apoptosis uses fragments of its body as a messenger to communicate with neighbors. It can ask to be extruded, as seen in the case of the intestinal epithelium, and, to this end, relies on help from junctional cell–cell communication structures [201]. It sends “find-me” messages to be recognized and “eat-me” signals to be engulfed by phagocytosis [62,459]. Even if it serves the purpose of finalizing cell death itself, the process of phagocytosis can be considered as a last intercellular communication of the dying cell with the cellular world [460]. Phagocytosis is not a mere waste collection system, it is actively involved in the apoptotic cell death process itself [461,462,463], although the mechanisms behind this are unknown. Here, a major difference between the modes of cell death is that, unlike apoptosis, cell death modes that lead to a necrotic outcome will communicate with their microenvironment via inflammation. While many research laboratories are looking into this, the call to action from dying cells towards intercellular communication molecules and structures still deserves greater attention. We hope that the examples used here will convey the importance of this aspect of cell death, linking it to cell communication.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Vaux, D.L.; Korsmeyer, S.J. Cell death in development. Cell 1999, 96, 245–254. [Google Scholar] [CrossRef]
  2. Kerr, J.F. History of the events leading to the formulation of the apoptosis concept. Toxicology 2002, 181–182, 471–474. [Google Scholar] [CrossRef] [PubMed]
  3. Kroemer, G.; Galluzzi, L.; Vandenabeele, P.; Abrams, J.; Alnemri, E.S.; Baehrecke, E.H.; Blagosklonny, M.V.; El-Deiry, W.S.; Golstein, P.; Green, D.R.; et al. Classification of cell death: Recommendations of the Nomenclature Committee on Cell Death 2009. Cell Death Differ. 2009, 16, 3–11. [Google Scholar] [CrossRef] [PubMed]
  4. Tang, H.L.; Yuen, K.L.; Tang, H.M.; Fung, M.C. Reversibility of apoptosis in cancer cells. Br. J. Cancer 2009, 100, 118–122. [Google Scholar] [CrossRef] [PubMed]
  5. Tang, H.L.; Tang, H.M.; Mak, K.H.; Hu, S.; Wang, S.S.; Wong, K.M.; Wong, C.S.; Wu, H.Y.; Law, H.T.; Liu, K.; et al. Cell survival, DNA damage, and oncogenic transformation after a transient and reversible apoptotic response. Mol. Biol. Cell 2012, 23, 2240–2252. [Google Scholar] [CrossRef] [PubMed]
  6. Zakharov, I.I.; Savitskaya, M.A.; Onishchenko, G.E. The Problem of Apoptotic Processes Reversibility. Biochemistry 2020, 85, 1145–1158. [Google Scholar] [CrossRef] [PubMed]
  7. Mohammed, R.N.; Khosravi, M.; Rahman, H.S.; Adili, A.; Kamali, N.; Soloshenkov, P.P.; Thangavelu, L.; Saeedi, H.; Shomali, N.; Tamjidifar, R.; et al. Anastasis: Cell recovery mechanisms and potential role in cancer. Cell Commun. Signal. 2022, 20, 81. [Google Scholar] [CrossRef] [PubMed]
  8. Zaitceva, V.; Kopeina, G.S.; Zhivotovsky, B. Anastasis: Return Journey from Cell Death. Cancers 2021, 13, 3671. [Google Scholar] [CrossRef]
  9. Chakraborty, S.; Mir, K.B.; Seligson, N.D.; Nayak, D.; Kumar, R.; Goswami, A. Integration of EMT and cellular survival instincts in reprogramming of programmed cell death to anastasis. Cancer Metastasis Rev. 2020, 39, 553–566. [Google Scholar] [CrossRef]
  10. Gong, Y.N.; Crawford, J.C.; Heckmann, B.L.; Green, D.R. To the edge of cell death and back. FEBS J. 2019, 286, 430–440. [Google Scholar] [CrossRef]
  11. Sun, G. Death and survival from executioner caspase activation. Semin. Cell Dev. Biol. 2024, 156, 66–73. [Google Scholar] [CrossRef]
  12. Lovric, M.M.; Hawkins, C.J. TRAIL treatment provokes mutations in surviving cells. Oncogene 2010, 29, 5048–5060. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, X.; He, Y.; Li, F.; Huang, Q.; Kato, T.A.; Hall, R.P.; Li, C.Y. Caspase-3 promotes genetic instability and carcinogenesis. Mol. Cell 2015, 58, 284–296. [Google Scholar] [CrossRef] [PubMed]
  14. Ichim, G.; Lopez, J.; Ahmed, S.U.; Muthalagu, N.; Giampazolias, E.; Delgado, M.E.; Haller, M.; Riley, J.S.; Mason, S.M.; Athineos, D.; et al. Limited mitochondrial permeabilization causes DNA damage and genomic instability in the absence of cell death. Mol. Cell 2015, 57, 860–872. [Google Scholar] [CrossRef] [PubMed]
  15. Sun, G.; Guzman, E.; Balasanyan, V.; Conner, C.M.; Wong, K.; Zhou, H.R.; Kosik, K.S.; Montell, D.J. A molecular signature for anastasis, recovery from the brink of apoptotic cell death. J. Cell Biol. 2017, 216, 3355–3368. [Google Scholar] [CrossRef] [PubMed]
  16. Nano, M.; Montell, D.J. Apoptotic signaling: Beyond cell death. Semin. Cell Dev. Biol. 2024, 156, 22–34. [Google Scholar] [CrossRef] [PubMed]
  17. Boya, P.; Kroemer, G. Lysosomal membrane permeabilization in cell death. Oncogene 2008, 27, 6434–6451. [Google Scholar] [CrossRef] [PubMed]
  18. Aits, S.; Jaattela, M. Lysosomal cell death at a glance. J. Cell Sci. 2013, 126 Pt 9, 1905–1912. [Google Scholar] [CrossRef]
  19. Battistelli, M.; Falcieri, E. Apoptotic Bodies: Particular Extracellular Vesicles Involved in Intercellular Communication. Biology 2020, 9, 21. [Google Scholar] [CrossRef]
  20. Kakarla, R.; Hur, J.; Kim, Y.J.; Kim, J.; Chwae, Y.J. Apoptotic cell-derived exosomes: Messages from dying cells. Exp. Mol. Med. 2020, 52, 1–6. [Google Scholar] [CrossRef]
  21. Zhang, M.; Lin, Y.; Chen, R.; Yu, H.; Li, Y.; Chen, M.; Dou, C.; Yin, P.; Zhang, L.; Tang, P. Ghost messages: Cell death signals spread. Cell Commun. Signal. 2023, 21, 6. [Google Scholar] [CrossRef]
  22. Golstein, P.; Kroemer, G. Cell death by necrosis: Towards a molecular definition. Trends Biochem. Sci. 2007, 32, 37–43. [Google Scholar] [CrossRef]
  23. Festjens, N.; Vanden Berghe, T.; Vandenabeele, P. Necrosis, a well-orchestrated form of cell demise: Signalling cascades, important mediators and concomitant immune response. Biochim. Biophys. Acta 2006, 1757, 1371–1387. [Google Scholar] [CrossRef]
  24. Silva, M.T.; do Vale, A.; dos Santos, N.M. Secondary necrosis in multicellular animals: An outcome of apoptosis with pathogenic implications. Apoptosis 2008, 13, 463–482. [Google Scholar] [CrossRef]
  25. Wyllie, A.H.; Kerr, J.F.; Currie, A.R. Cell death: The significance of apoptosis. Int. Rev. Cytol. 1980, 68, 251–306. [Google Scholar] [CrossRef] [PubMed]
  26. Gaipl, U.S.; Franz, S.; Voll, R.E.; Sheriff, A.; Kalden, J.R.; Herrmann, M. Defects in the disposal of dying cells lead to autoimmunity. Curr. Rheumatol. Rep. 2004, 6, 401–407. [Google Scholar] [CrossRef] [PubMed]
  27. Lauber, K.; Blumenthal, S.G.; Waibel, M.; Wesselborg, S. Clearance of apoptotic cells: Getting rid of the corpses. Mol. Cell 2004, 14, 277–287. [Google Scholar] [CrossRef]
  28. de Castro, M.A.; Bunt, G.; Wouters, F.S. Cathepsin B launches an apoptotic exit effort upon cell death-associated disruption of lysosomes. Cell Death Discov. 2016, 2, 16012. [Google Scholar] [CrossRef]
  29. Vercammen, D.; Beyaert, R.; Denecker, G.; Goossens, V.; Van Loo, G.; Declercq, W.; Grooten, J.; Fiers, W.; Vandenabeele, P. Inhibition of caspases increases the sensitivity of L929 cells to necrosis mediated by tumor necrosis factor. J. Exp. Med. 1998, 187, 1477–1485. [Google Scholar] [CrossRef] [PubMed]
  30. Holler, N.; Zaru, R.; Micheau, O.; Thome, M.; Attinger, A.; Valitutti, S.; Bodmer, J.L.; Schneider, P.; Seed, B.; Tschopp, J. Fas triggers an alternative, caspase-8-independent cell death pathway using the kinase RIP as effector molecule. Nat. Immunol. 2000, 1, 489–495. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, D.W.; Shao, J.; Lin, J.; Zhang, N.; Lu, B.J.; Lin, S.C.; Dong, M.Q.; Han, J. RIP3, an energy metabolism regulator that switches TNF-induced cell death from apoptosis to necrosis. Science 2009, 325, 332–336. [Google Scholar] [CrossRef] [PubMed]
  32. Kaczmarek, A.; Vandenabeele, P.; Krysko, D.V. Necroptosis: The release of damage-associated molecular patterns and its physiological relevance. Immunity 2013, 38, 209–223. [Google Scholar] [CrossRef] [PubMed]
  33. Newton, K.; Manning, G. Necroptosis and Inflammation. Annu. Rev. Biochem. 2016, 85, 743–763. [Google Scholar] [CrossRef] [PubMed]
  34. Chan, F.K.; Luz, N.F.; Moriwaki, K. Programmed necrosis in the cross talk of cell death and inflammation. Annu. Rev. Immunol. 2015, 33, 79–106. [Google Scholar] [CrossRef] [PubMed]
  35. Pasparakis, M.; Vandenabeele, P. Necroptosis and its role in inflammation. Nature 2015, 517, 311–320. [Google Scholar] [CrossRef] [PubMed]
  36. Nailwal, H.; Chan, F.K. Necroptosis in anti-viral inflammation. Cell Death Differ. 2019, 26, 4–13. [Google Scholar] [CrossRef] [PubMed]
  37. Kovacs, S.B.; Miao, E.A. Gasdermins: Effectors of Pyroptosis. Trends Cell Biol. 2017, 27, 673–684. [Google Scholar] [CrossRef]
  38. Shi, J.; Gao, W.; Shao, F. Pyroptosis: Gasdermin-Mediated Programmed Necrotic Cell Death. Trends Biochem. Sci. 2017, 42, 245–254. [Google Scholar] [CrossRef]
  39. Cookson, B.T.; Brennan, M.A. Pro-inflammatory programmed cell death. Trends Microbiol. 2001, 9, 113–114. [Google Scholar] [CrossRef]
  40. Pilla, D.M.; Hagar, J.A.; Haldar, A.K.; Mason, A.K.; Degrandi, D.; Pfeffer, K.; Ernst, R.K.; Yamamoto, M.; Miao, E.A.; Coers, J. Guanylate binding proteins promote caspase-11-dependent pyroptosis in response to cytoplasmic LPS. Proc. Natl. Acad. Sci. USA 2014, 111, 6046–6051. [Google Scholar] [CrossRef]
  41. Shi, J.; Zhao, Y.; Wang, K.; Shi, X.; Wang, Y.; Huang, H.; Zhuang, Y.; Cai, T.; Wang, F.; Shao, F. Cleavage of GSDMD by inflammatory caspases determines pyroptotic cell death. Nature 2015, 526, 660–665. [Google Scholar] [CrossRef]
  42. Jorgensen, I.; Miao, E.A. Pyroptotic cell death defends against intracellular pathogens. Immunol. Rev. 2015, 265, 130–142. [Google Scholar] [CrossRef] [PubMed]
  43. Miao, E.A.; Leaf, I.A.; Treuting, P.M.; Mao, D.P.; Dors, M.; Sarkar, A.; Warren, S.E.; Wewers, M.D.; Aderem, A. Caspase-1-induced pyroptosis is an innate immune effector mechanism against intracellular bacteria. Nat. Immunol. 2010, 11, 1136–1142. [Google Scholar] [CrossRef] [PubMed]
  44. Aachoui, Y.; Leaf, I.A.; Hagar, J.A.; Fontana, M.F.; Campos, C.G.; Zak, D.E.; Tan, M.H.; Cotter, P.A.; Vance, R.E.; Aderem, A.; et al. Caspase-11 protects against bacteria that escape the vacuole. Science 2013, 339, 975–978. [Google Scholar] [CrossRef] [PubMed]
  45. 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]
  46. Sborgi, L.; Ruhl, S.; Mulvihill, E.; Pipercevic, J.; Heilig, R.; Stahlberg, H.; Farady, C.J.; Muller, D.J.; Broz, P.; Hiller, S. GSDMD membrane pore formation constitutes the mechanism of pyroptotic cell death. EMBO J. 2016, 35, 1766–1778. [Google Scholar] [CrossRef] [PubMed]
  47. Ding, J.; Wang, K.; Liu, W.; She, Y.; Sun, Q.; Shi, J.; Sun, H.; Wang, D.C.; Shao, F. Pore-forming activity and structural autoinhibition of the gasdermin family. Nature 2016, 535, 111–116. [Google Scholar] [CrossRef]
  48. Liu, X.; Zhang, Z.; Ruan, J.; Pan, Y.; Magupalli, V.G.; Wu, H.; Lieberman, J. Inflammasome-activated gasdermin D causes pyroptosis by forming membrane pores. Nature 2016, 535, 153–158. [Google Scholar] [CrossRef]
  49. Man, S.M.; Kanneganti, T.D. Converging roles of caspases in inflammasome activation, cell death and innate immunity. Nat. Rev. Immunol. 2016, 16, 7–21. [Google Scholar] [CrossRef]
  50. Lockshin, R.A.; Williams, C.M. Programmed cell death. IV. The influence of drugs on the breakdown of the intersegmental muscles of silkmoths. J. Insect. Physiol. 1965, 11, 803–809. [Google Scholar] [CrossRef]
  51. Tata, J.R. Requirement for RNA and protein synthesis for induced regression of the tadpole tail in organ culture. Dev. Biol. 1966, 13, 77–94. [Google Scholar] [CrossRef] [PubMed]
  52. Kerr, J.F.; Wyllie, A.H.; Currie, A.R. Apoptosis: A basic biological phenomenon with wide-ranging implications in tissue kinetics. Br. J. Cancer 1972, 26, 239–257. [Google Scholar] [CrossRef] [PubMed]
  53. Nossing, C.; Ryan, K.M. 50 years on and still very much alive: Apoptosis: A basic biological phenomenon with wide-ranging implications in tissue kinetics. Br. J. Cancer 2023, 128, 426–431. [Google Scholar] [CrossRef] [PubMed]
  54. Horvitz, H.R. Genetic control of programmed cell death in the nematode Caenorhabditis elegans. Cancer Res. 1999, 59 (Suppl. S7), 1701s–1706s. [Google Scholar] [PubMed]
  55. Abrams, J.M.; White, K.; Fessler, L.I.; Steller, H. Programmed cell death during Drosophila embryogenesis. Development 1993, 117, 29–43. [Google Scholar] [CrossRef] [PubMed]
  56. Xu, D.; Woodfield, S.E.; Lee, T.V.; Fan, Y.; Antonio, C.; Bergmann, A. Genetic control of programmed cell death (apoptosis) in Drosophila. Fly 2009, 3, 78–90. [Google Scholar] [CrossRef]
  57. Ranger, A.M.; Malynn, B.A.; Korsmeyer, S.J. Mouse models of cell death. Nat. Genet. 2001, 28, 113–118. [Google Scholar] [CrossRef] [PubMed]
  58. Joza, N.; Kroemer, G.; Penninger, J.M. Genetic analysis of the mammalian cell death machinery. Trends Genet. 2002, 18, 142–149. [Google Scholar] [CrossRef]
  59. Woo, M.; Hakem, R.; Mak, T.W. Executionary pathway for apoptosis: Lessons from mutant mice. Cell Res. 2000, 10, 267–278. [Google Scholar] [CrossRef]
  60. Fink, S.L.; Cookson, B.T. Apoptosis, pyroptosis, and necrosis: Mechanistic description of dead and dying eukaryotic cells. Infect. Immun. 2005, 73, 1907–1916. [Google Scholar] [CrossRef]
  61. Davidovich, P.; Kearney, C.J.; Martin, S.J. Inflammatory outcomes of apoptosis, necrosis and necroptosis. Biol. Chem. 2014, 395, 1163–1171. [Google Scholar] [CrossRef] [PubMed]
  62. Nagata, S. Apoptosis and Clearance of Apoptotic Cells. Annu. Rev. Immunol. 2018, 36, 489–517. [Google Scholar] [CrossRef] [PubMed]
  63. 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]
  64. Snyder, A.G.; Oberst, A. The Antisocial Network: Cross Talk Between Cell Death Programs in Host Defense. Annu. Rev. Immunol. 2021, 39, 77–101. [Google Scholar] [CrossRef] [PubMed]
  65. Bedoui, S.; Herold, M.J.; Strasser, A. Emerging connectivity of programmed cell death pathways and its physiological implications. Nat. Rev. Mol. Cell Biol. 2020, 21, 678–695. [Google Scholar] [CrossRef]
  66. 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]
  67. Mizushima, N. A brief history of autophagy from cell biology to physiology and disease. Nat. Cell Biol. 2018, 20, 521–527. [Google Scholar] [CrossRef]
  68. Klionsky, D.J. Autophagy revisited: A conversation with Christian de Duve. Autophagy 2008, 4, 740–743. [Google Scholar] [CrossRef]
  69. Mizushima, N.; Komatsu, M. Autophagy: Renovation of cells and tissues. Cell 2011, 147, 728–741. [Google Scholar] [CrossRef]
  70. Vargas, J.N.S.; Hamasaki, M.; Kawabata, T.; Youle, R.J.; Yoshimori, T. The mechanisms and roles of selective autophagy in mammals. Nat. Rev. Mol. Cell Biol. 2023, 24, 167–185. [Google Scholar] [CrossRef]
  71. Wang, L.; Klionsky, D.J.; Shen, H.M. The emerging mechanisms and functions of microautophagy. Nat. Rev. Mol. Cell Biol. 2023, 24, 186–203. [Google Scholar] [CrossRef] [PubMed]
  72. Kim, K.H.; Lee, M.S. Autophagy—A key player in cellular and body metabolism. Nat. Rev. Endocrinol. 2014, 10, 322–337. [Google Scholar] [CrossRef] [PubMed]
  73. Maiuri, M.C.; Zalckvar, E.; Kimchi, A.; Kroemer, G. Self-eating and self-killing: Crosstalk between autophagy and apoptosis. Nat. Rev. Mol. Cell Biol. 2007, 8, 741–752. [Google Scholar] [CrossRef] [PubMed]
  74. Ploumi, C.; Papandreou, M.E.; Tavernarakis, N. The complex interplay between autophagy and cell death pathways. Biochem. J. 2022, 479, 75–90. [Google Scholar] [CrossRef] [PubMed]
  75. Denton, D.; Kumar, S. Autophagy-dependent cell death. Cell Death Differ. 2019, 26, 605–616. [Google Scholar] [CrossRef] [PubMed]
  76. Loos, B.; Engelbrecht, A.M.; Lockshin, R.A.; Klionsky, D.J.; Zakeri, Z. The variability of autophagy and cell death susceptibility: Unanswered questions. Autophagy 2013, 9, 1270–1285. [Google Scholar] [CrossRef] [PubMed]
  77. Mahapatra, K.K.; Mishra, S.R.; Behera, B.P.; Patil, S.; Gewirtz, D.A.; Bhutia, S.K. The lysosome as an imperative regulator of autophagy and cell death. Cell Mol. Life Sci. 2021, 78, 7435–7449. [Google Scholar] [CrossRef]
  78. Johansson, A.C.; Appelqvist, H.; Nilsson, C.; Kagedal, K.; Roberg, K.; Ollinger, K. Regulation of apoptosis-associated lysosomal membrane permeabilization. Apoptosis 2010, 15, 527–540. [Google Scholar] [CrossRef]
  79. Skowyra, M.L.; Schlesinger, P.H.; Naismith, T.V.; Hanson, P.I. Triggered recruitment of ESCRT machinery promotes endolysosomal repair. Science 2018, 360, eaar5078. [Google Scholar] [CrossRef]
  80. Radulovic, M.; Schink, K.O.; Wenzel, E.M.; Nahse, V.; Bongiovanni, A.; Lafont, F.; Stenmark, H. ESCRT-mediated lysosome repair precedes lysophagy and promotes cell survival. EMBO J. 2018, 37, e99753. [Google Scholar] [CrossRef]
  81. Ono, K.; Kim, S.O.; Han, J. Susceptibility of lysosomes to rupture is a determinant for plasma membrane disruption in tumor necrosis factor alpha-induced cell death. Mol. Cell Biol. 2003, 23, 665–676. [Google Scholar] [CrossRef]
  82. Galluzzi, L.; Bravo-San Pedro, J.M.; Kroemer, G. Organelle-specific initiation of cell death. Nat. Cell Biol. 2014, 16, 728–736. [Google Scholar] [CrossRef] [PubMed]
  83. Joshi, G.N.; Knecht, D.A. Silica phagocytosis causes apoptosis and necrosis by different temporal and molecular pathways in alveolar macrophages. Apoptosis 2013, 18, 271–285. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, F.; Gomez-Sintes, R.; Boya, P. Lysosomal membrane permeabilization and cell death. Traffic 2018, 19, 918–931. [Google Scholar] [CrossRef] [PubMed]
  85. Eaton, J.W.; Qian, M. Molecular bases of cellular iron toxicity. Free Radic. Biol. Med. 2002, 32, 833–840. [Google Scholar] [CrossRef] [PubMed]
  86. Borner, C.; Monney, L. Apoptosis without caspases: An inefficient molecular guillotine? Cell Death Differ. 1999, 6, 497–507. [Google Scholar] [CrossRef]
  87. Mrschtik, M.; Ryan, K.M. Lysosomal proteins in cell death and autophagy. FEBS J. 2015, 282, 1858–1870. [Google Scholar] [CrossRef] [PubMed]
  88. Counis, M.F.; Torriglia, A. DNases and apoptosis. Biochem. Cell Biol. 2000, 78, 405–414. [Google Scholar] [CrossRef]
  89. Keyel, P.A. Dnases in health and disease. Dev. Biol. 2017, 429, 1–11. [Google Scholar] [CrossRef]
  90. Daugas, E.; Susin, S.A.; Zamzami, N.; Ferri, K.F.; Irinopoulou, T.; Larochette, N.; Prevost, M.C.; Leber, B.; Andrews, D.; Penninger, J.; et al. Mitochondrio-nuclear translocation of AIF in apoptosis and necrosis. FASEB J. 2000, 14, 729–739. [Google Scholar] [CrossRef]
  91. Li, L.Y.; Luo, X.; Wang, X. Endonuclease G is an apoptotic DNase when released from mitochondria. Nature 2001, 412, 95–99. [Google Scholar] [CrossRef] [PubMed]
  92. Martin, S.J.; Green, D.R. Protease activation during apoptosis: Death by a thousand cuts? Cell 1995, 82, 349–352. [Google Scholar] [CrossRef] [PubMed]
  93. He, B.; Lu, N.; Zhou, Z. Cellular and nuclear degradation during apoptosis. Curr. Opin. Cell Biol. 2009, 21, 900–912. [Google Scholar] [CrossRef]
  94. Taylor, R.C.; Cullen, S.P.; Martin, S.J. Apoptosis: Controlled demolition at the cellular level. Nat. Rev. Mol. Cell Biol. 2008, 9, 231–241. [Google Scholar] [CrossRef] [PubMed]
  95. Adrain, C.; Brumatti, G.; Martin, S.J. Apoptosomes: Protease activation platforms to die from. Trends Biochem. Sci. 2006, 31, 243–247. [Google Scholar] [CrossRef]
  96. Dorstyn, L.; Akey, C.W.; Kumar, S. New insights into apoptosome structure and function. Cell Death Differ. 2018, 25, 1194–1208. [Google Scholar] [CrossRef]
  97. Williams, J.R.; Little, J.B.; Shipley, W.U. Association of mammalian cell death with a specific endonucleolytic degradation of DNA. Nature 1974, 252, 754–755. [Google Scholar] [CrossRef] [PubMed]
  98. Luthi, A.U.; Martin, S.J. The CASBAH: A searchable database of caspase substrates. Cell Death Differ. 2007, 14, 641–650. [Google Scholar] [CrossRef]
  99. Monier, B.; Suzanne, M. Orchestration of Force Generation and Nuclear Collapse in Apoptotic Cells. Int. J. Mol. Sci. 2021, 22, 10257. [Google Scholar] [CrossRef]
  100. Stewart, C.L.; Roux, K.J.; Burke, B. Blurring the boundary: The nuclear envelope extends its reach. Science 2007, 318, 1408–1412. [Google Scholar] [CrossRef]
  101. Burke, B. Lamins and apoptosis: A two-way street? J. Cell Biol. 2001, 153, F5–F7. [Google Scholar] [CrossRef] [PubMed]
  102. Rao, L.; Perez, D.; White, E. Lamin proteolysis facilitates nuclear events during apoptosis. J. Cell Biol. 1996, 135 Pt 1, 1441–1455. [Google Scholar] [CrossRef] [PubMed]
  103. Walker, A.; Ward, C.; Sheldrake, T.A.; Dransfield, I.; Rossi, A.G.; Pryde, J.G.; Haslett, C. Golgi fragmentation during Fas-mediated apoptosis is associated with the rapid loss of GM130. Biochem. Biophys. Res. Commun. 2004, 316, 6–11. [Google Scholar] [CrossRef] [PubMed]
  104. Lane, J.D.; Lucocq, J.; Pryde, J.; Barr, F.A.; Woodman, P.G.; Allan, V.J.; Lowe, M. Caspase-mediated cleavage of the stacking protein GRASP65 is required for Golgi fragmentation during apoptosis. J. Cell Biol. 2002, 156, 495–509. [Google Scholar] [CrossRef] [PubMed]
  105. Robertson, J.D.; Orrenius, S.; Zhivotovsky, B. Review: Nuclear events in apoptosis. J. Struct. Biol. 2000, 129, 346–358. [Google Scholar] [CrossRef] [PubMed]
  106. Doonan, F.; Cotter, T.G. Morphological assessment of apoptosis. Methods 2008, 44, 200–204. [Google Scholar] [CrossRef] [PubMed]
  107. Buendia, B.; Courvalin, J.C.; Collas, P. Dynamics of the nuclear envelope at mitosis and during apoptosis. Cell Mol. Life Sci. 2001, 58, 1781–1789. [Google Scholar] [CrossRef]
  108. Kutay, U.; Hetzer, M.W. Reorganization of the nuclear envelope during open mitosis. Curr. Opin. Cell Biol. 2008, 20, 669–677. [Google Scholar] [CrossRef]
  109. Ungricht, R.; Kutay, U. Mechanisms and functions of nuclear envelope remodelling. Nat. Rev. Mol. Cell Biol. 2017, 18, 229–245. [Google Scholar] [CrossRef]
  110. Shahin, V. Strategic disruption of nuclear pores structure, integrity and barrier for nuclear apoptosis. Semin. Cell Dev. Biol. 2017, 68, 85–90. [Google Scholar] [CrossRef]
  111. Youle, R.J.; van der Bliek, A.M. Mitochondrial fission, fusion, and stress. Science 2012, 337, 1062–1065. [Google Scholar] [CrossRef] [PubMed]
  112. Hoppins, S.; Nunnari, J. Cell Biology. Mitochondrial dynamics and apoptosis--the ER connection. Science 2012, 337, 1052–1054. [Google Scholar] [CrossRef] [PubMed]
  113. Machamer, C.E. Golgi disassembly in apoptosis: Cause or effect? Trends Cell Biol. 2003, 13, 279–281. [Google Scholar] [CrossRef] [PubMed]
  114. Gonatas, N.K.; Stieber, A.; Gonatas, J.O. Fragmentation of the Golgi apparatus in neurodegenerative diseases and cell death. J. Neurol. Sci. 2006, 246, 21–30. [Google Scholar] [CrossRef] [PubMed]
  115. Rao, R.V.; Ellerby, H.M.; Bredesen, D.E. Coupling endoplasmic reticulum stress to the cell death program. Cell Death Differ. 2004, 11, 372–380. [Google Scholar] [CrossRef]
  116. Suen, D.F.; Norris, K.L.; Youle, R.J. Mitochondrial dynamics and apoptosis. Genes Dev. 2008, 22, 1577–1590. [Google Scholar] [CrossRef] [PubMed]
  117. Elgass, K.; Pakay, J.; Ryan, M.T.; Palmer, C.S. Recent advances into the understanding of mitochondrial fission. Biochim. Biophys. Acta 2013, 1833, 150–161. [Google Scholar] [CrossRef]
  118. Arnoult, D. Mitochondrial fragmentation in apoptosis. Trends Cell Biol. 2007, 17, 6–12. [Google Scholar] [CrossRef]
  119. Martinou, J.C.; Youle, R.J. Which came first, the cytochrome c release or the mitochondrial fission? Cell Death Differ. 2006, 13, 1291–1295. [Google Scholar] [CrossRef]
  120. Bock, F.J.; Tait, S.W.G. Mitochondria as multifaceted regulators of cell death. Nat. Rev. Mol. Cell Biol. 2020, 21, 85–100. [Google Scholar] [CrossRef]
  121. Lane, J.D.; Allan, V.J.; Woodman, P.G. Active relocation of chromatin and endoplasmic reticulum into blebs in late apoptotic cells. J. Cell Sci. 2005, 118 Pt 17, 4059–4071. [Google Scholar] [CrossRef] [PubMed]
  122. Keller, H.; Rentsch, P.; Hagmann, J. Differences in cortical actin structure and dynamics document that different types of blebs are formed by distinct mechanisms. Exp. Cell Res. 2002, 277, 161–172. [Google Scholar] [CrossRef] [PubMed]
  123. Mills, J.C.; Stone, N.L.; Pittman, R.N. Extranuclear apoptosis. The role of the cytoplasm in the execution phase. J. Cell Biol. 1999, 146, 703–708. [Google Scholar] [CrossRef] [PubMed]
  124. Mills, J.C.; Stone, N.L.; Erhardt, J.; Pittman, R.N. Apoptotic membrane blebbing is regulated by myosin light chain phosphorylation. J. Cell Biol. 1998, 140, 627–636. [Google Scholar] [CrossRef] [PubMed]
  125. Charras, G.T.; Yarrow, J.C.; Horton, M.A.; Mahadevan, L.; Mitchison, T.J. Non-equilibration of hydrostatic pressure in blebbing cells. Nature 2005, 435, 365–369. [Google Scholar] [CrossRef] [PubMed]
  126. Croft, D.R.; Coleman, M.L.; Li, S.; Robertson, D.; Sullivan, T.; Stewart, C.L.; Olson, M.F. Actin-myosin-based contraction is responsible for apoptotic nuclear disintegration. J. Cell Biol. 2005, 168, 245–255. [Google Scholar] [CrossRef] [PubMed]
  127. Moss, D.K.; Betin, V.M.; Malesinski, S.D.; Lane, J.D. A novel role for microtubules in apoptotic chromatin dynamics and cellular fragmentation. J. Cell Sci. 2006, 119 Pt 11, 2362–2374. [Google Scholar] [CrossRef] [PubMed]
  128. Bozzaro, S. The past, present and future of Dictyostelium as a model system. Int. J. Dev. Biol. 2019, 63, 321–331. [Google Scholar] [CrossRef]
  129. Kandouz, M. Dying to communicate: Apoptotic functions of Eph/Ephrin proteins. Apoptosis 2018, 23, 265–289. [Google Scholar] [CrossRef]
  130. Bates, R.C.; Buret, A.; van Helden, D.F.; Horton, M.A.; Burns, G.F. Apoptosis induced by inhibition of intercellular contact. J. Cell Biol. 1994, 125, 403–415. [Google Scholar] [CrossRef]
  131. Krysko, D.V.; Leybaert, L.; Vandenabeele, P.; D’Herde, K. Gap junctions and the propagation of cell survival and cell death signals. Apoptosis 2005, 10, 459–469. [Google Scholar] [CrossRef]
  132. Carette, D.; Gilleron, J.; Chevallier, D.; Segretain, D.; Pointis, G. Connexin a check-point component of cell apoptosis in normal and physiopathological conditions. Biochimie 2014, 101, 1–9. [Google Scholar] [CrossRef] [PubMed]
  133. Gilleron, J.; Carette, D.; Segretain, D.; Pointis, G. Multiple and complex influences of connexins and pannexins on cell death. Biochim. Biophys. Acta Biomembr. 2018, 1860, 182–191. [Google Scholar] [CrossRef] [PubMed]
  134. Caruso, S.; Poon, I.K.H. Apoptotic Cell-Derived Extracellular Vesicles: More Than Just Debris. Front. Immunol. 2018, 9, 1486. [Google Scholar] [CrossRef] [PubMed]
  135. Sanwlani, R.; Gangoda, L. Role of Extracellular Vesicles in Cell Death and Inflammation. Cells 2021, 10, 2663. [Google Scholar] [CrossRef] [PubMed]
  136. Wang, X.; Gerdes, H.H. Transfer of mitochondria via tunneling nanotubes rescues apoptotic PC12 cells. Cell Death Differ. 2015, 22, 1181–1191. [Google Scholar] [CrossRef]
  137. Zheng, F.; Luo, Z.; Lin, X.; Wang, W.; Aschner, M.; Cai, P.; Wang, Y.L.; Shao, W.; Yu, G.; Guo, Z.; et al. Intercellular transfer of mitochondria via tunneling nanotubes protects against cobalt nanoparticle-induced neurotoxicity and mitochondrial damage. Nanotoxicology 2021, 15, 1358–1379. [Google Scholar] [CrossRef] [PubMed]
  138. Guo, R.; Davis, D.; Fang, Y. Intercellular transfer of mitochondria rescues virus-induced cell death but facilitates cell-to-cell spreading of porcine reproductive and respiratory syndrome virus. Virology 2018, 517, 122–134. [Google Scholar] [CrossRef] [PubMed]
  139. Luchetti, F.; Canonico, B.; Arcangeletti, M.; Guescini, M.; Cesarini, E.; Stocchi, V.; Degli Esposti, M.; Papa, S. Fas signalling promotes intercellular communication in T cells. PLoS ONE 2012, 7, e35766. [Google Scholar] [CrossRef]
  140. Bittins, M.; Wang, X. TNT-Induced Phagocytosis: Tunneling Nanotubes Mediate the Transfer of Pro-Phagocytic Signals From Apoptotic to Viable Cells. J. Cell. Physiol. 2017, 232, 2271–2279. [Google Scholar] [CrossRef]
  141. Lu, S.; Sun, Z.; Liu, L.; Li, P.; Li, B.; Li, W.; Wu, Z.; Zhao, M.; Liu, W.; Wang, Y.; et al. Tumor-Derived Exosomes Regulate Apoptosis of CD45(+)EpCAM(+) Cells in Lung Cancer. Front. Immunol. 2022, 13, 903882. [Google Scholar] [CrossRef] [PubMed]
  142. Fuchs, Y.; Steller, H. Programmed cell death in animal development and disease. Cell 2011, 147, 742–758. [Google Scholar] [CrossRef] [PubMed]
  143. Voss, A.K.; Strasser, A. The essentials of developmental apoptosis. F1000Research 2020, 9. [Google Scholar] [CrossRef] [PubMed]
  144. Wong, F.K.; Marin, O. Developmental Cell Death in the Cerebral Cortex. Annu. Rev. Cell Dev. Biol. 2019, 35, 523–542. [Google Scholar] [CrossRef] [PubMed]
  145. Southwell, D.G.; Paredes, M.F.; Galvao, R.P.; Jones, D.L.; Froemke, R.C.; Sebe, J.Y.; Alfaro-Cervello, C.; Tang, Y.; Garcia-Verdugo, J.M.; Rubenstein, J.L.; et al. Intrinsically determined cell death of developing cortical interneurons. Nature 2012, 491, 109–113. [Google Scholar] [CrossRef]
  146. Doseff, A.I. Apoptosis: The sculptor of development. Stem Cells Dev. 2004, 13, 473–483. [Google Scholar] [CrossRef] [PubMed]
  147. Boehm, I. Apoptosis in physiological and pathological skin: Implications for therapy. Curr. Mol. Med. 2006, 6, 375–394. [Google Scholar] [CrossRef] [PubMed]
  148. Anderton, H.; Alqudah, S. Cell death in skin function, inflammation, and disease. Biochem. J. 2022, 479, 1621–1651. [Google Scholar] [CrossRef]
  149. Tower, J. Programmed cell death in aging. Ageing Res. Rev. 2015, 23 Pt A, 90–100. [Google Scholar] [CrossRef]
  150. Hardy, K.; Handyside, A.H. Metabolism and cell allocation during parthenogenetic preimplantation mouse development. Mol. Reprod. Dev. 1996, 43, 313–322. [Google Scholar] [CrossRef]
  151. Fleming, T.P.; Sheth, B.; Fesenko, I. Cell adhesion in the preimplantation mammalian embryo and its role in trophectoderm differentiation and blastocyst morphogenesis. Front. Biosci. 2001, 6, D1000-7. [Google Scholar] [CrossRef]
  152. Bloor, D.J.; Metcalfe, A.D.; Rutherford, A.; Brison, D.R.; Kimber, S.J. Expression of cell adhesion molecules during human preimplantation embryo development. Mol. Hum. Reprod. 2002, 8, 237–245. [Google Scholar] [CrossRef]
  153. Brison, D.R.; Sturmey, R.G.; Leese, H.J. Metabolic heterogeneity during preimplantation development: The missing link? Hum. Reprod. Update 2014, 20, 632–640. [Google Scholar] [CrossRef] [PubMed]
  154. Venturas, M.; Shah, J.S.; Yang, X.; Sanchez, T.H.; Conway, W.; Sakkas, D.; Needleman, D.J. Metabolic state of human blastocysts measured by fluorescence lifetime imaging microscopy. Hum. Reprod. 2022, 37, 411–427. [Google Scholar] [CrossRef] [PubMed]
  155. Leese, H.J. Quiet please, do not disturb: A hypothesis of embryo metabolism and viability. Bioessays 2002, 24, 845–849. [Google Scholar] [CrossRef] [PubMed]
  156. Baumann, C.G.; Morris, D.G.; Sreenan, J.M.; Leese, H.J. The quiet embryo hypothesis: Molecular characteristics favoring viability. Mol. Reprod. Dev. 2007, 74, 1345–1353. [Google Scholar] [CrossRef] [PubMed]
  157. Leese, H.J. Metabolism of the preimplantation embryo: 40 years on. Reproduction 2012, 143, 417–427. [Google Scholar] [CrossRef] [PubMed]
  158. Byrne, A.T.; Southgate, J.; Brison, D.R.; Leese, H.J. Analysis of apoptosis in the preimplantation bovine embryo using TUNEL. J. Reprod. Fertil. 1999, 117, 97–105. [Google Scholar] [CrossRef] [PubMed]
  159. Gjorret, J.O.; Knijn, H.M.; Dieleman, S.J.; Avery, B.; Larsson, L.I.; Maddox-Hyttel, P. Chronology of apoptosis in bovine embryos produced in vivo and in vitro. Biol. Reprod. 2003, 69, 1193–1200. [Google Scholar] [CrossRef]
  160. Brison, D.R.; Schultz, R.M. Apoptosis during mouse blastocyst formation: Evidence for a role for survival factors including transforming growth factor alpha. Biol. Reprod. 1997, 56, 1088–1096. [Google Scholar] [CrossRef]
  161. Hardy, K. Cell death in the mammalian blastocyst. Mol. Hum. Reprod. 1997, 3, 919–925. [Google Scholar] [CrossRef]
  162. Moura, M.T.; Latorraca, L.B.; Paula-Lopes, F.F. Contextualizing Autophagy during Gametogenesis and Preimplantation Embryonic Development. Int. J. Mol. Sci. 2021, 22, 6313. [Google Scholar] [CrossRef] [PubMed]
  163. Escobar, M.L.; Echeverria, O.M.; Ortiz, R.; Vazquez-Nin, G.H. Combined apoptosis and autophagy, the process that eliminates the oocytes of atretic follicles in immature rats. Apoptosis 2008, 13, 1253–1266. [Google Scholar] [CrossRef] [PubMed]
  164. Song, W.H.; Yi, Y.J.; Sutovsky, M.; Meyers, S.; Sutovsky, P. Autophagy and ubiquitin-proteasome system contribute to sperm mitophagy after mammalian fertilization. Proc. Natl. Acad. Sci. USA 2016, 113, E5261–E5270. [Google Scholar] [CrossRef] [PubMed]
  165. Tsukamoto, S.; Kuma, A.; Murakami, M.; Kishi, C.; Yamamoto, A.; Mizushima, N. Autophagy is essential for preimplantation development of mouse embryos. Science 2008, 321, 117–120. [Google Scholar] [CrossRef] [PubMed]
  166. Tsukamoto, S.; Kuma, A.; Mizushima, N. The role of autophagy during the oocyte-to-embryo transition. Autophagy 2008, 4, 1076–1078. [Google Scholar] [CrossRef] [PubMed]
  167. Al Rawi, S.; Louvet-Vallee, S.; Djeddi, A.; Sachse, M.; Culetto, E.; Hajjar, C.; Boyd, L.; Legouis, R.; Galy, V. Postfertilization autophagy of sperm organelles prevents paternal mitochondrial DNA transmission. Science 2011, 334, 1144–1147. [Google Scholar] [CrossRef] [PubMed]
  168. Sato, M.; Sato, K. Degradation of paternal mitochondria by fertilization-triggered autophagy in C. elegans embryos. Science 2011, 334, 1141–1144. [Google Scholar] [CrossRef] [PubMed]
  169. Politi, Y.; Gal, L.; Kalifa, Y.; Ravid, L.; Elazar, Z.; Arama, E. Paternal mitochondrial destruction after fertilization is mediated by a common endocytic and autophagic pathway in Drosophila. Dev. Cell 2014, 29, 305–320. [Google Scholar] [CrossRef]
  170. Allen, E.A.; Baehrecke, E.H. Autophagy in animal development. Cell Death Differ. 2020, 27, 903–918. [Google Scholar] [CrossRef]
  171. Qu, X.; Zou, Z.; Sun, Q.; Luby-Phelps, K.; Cheng, P.; Hogan, R.N.; Gilpin, C.; Levine, B. Autophagy gene-dependent clearance of apoptotic cells during embryonic development. Cell 2007, 128, 931–946. [Google Scholar] [CrossRef] [PubMed]
  172. Qu, X.; Yu, J.; Bhagat, G.; Furuya, N.; Hibshoosh, H.; Troxel, A.; Rosen, J.; Eskelinen, E.L.; Mizushima, N.; Ohsumi, Y.; et al. Promotion of tumorigenesis by heterozygous disruption of the beclin 1 autophagy gene. J. Clin. Investig. 2003, 112, 1809–1820. [Google Scholar] [CrossRef] [PubMed]
  173. Yue, Z.; Jin, S.; Yang, C.; Levine, A.J.; Heintz, N. Beclin 1, an autophagy gene essential for early embryonic development, is a haploinsufficient tumor suppressor. Proc. Natl. Acad. Sci. USA 2003, 100, 15077–15082. [Google Scholar] [CrossRef] [PubMed]
  174. Hara, T.; Nakamura, K.; Matsui, M.; Yamamoto, A.; Nakahara, Y.; Suzuki-Migishima, R.; Yokoyama, M.; Mishima, K.; Saito, I.; Okano, H.; et al. Suppression of basal autophagy in neural cells causes neurodegenerative disease in mice. Nature 2006, 441, 885–889. [Google Scholar] [CrossRef] [PubMed]
  175. Liang, C.C.; Wang, C.; Peng, X.; Gan, B.; Guan, J.L. Neural-specific deletion of FIP200 leads to cerebellar degeneration caused by increased neuronal death and axon degeneration. J. Biol. Chem. 2010, 285, 3499–3509. [Google Scholar] [CrossRef] [PubMed]
  176. Le Douarin, N.M.; Dupin, E. The “beginnings” of the neural crest. Dev. Biol. 2018, 444 (Suppl. S1), S3–S13. [Google Scholar] [CrossRef] [PubMed]
  177. Dupin, E.; Calloni, G.W.; Coelho-Aguiar, J.M.; Le Douarin, N.M. The issue of the multipotency of the neural crest cells. Dev. Biol. 2018, 444 (Suppl. S1), S47–S59. [Google Scholar] [CrossRef] [PubMed]
  178. Szabo, A.; Mayor, R. Mechanisms of Neural Crest Migration. Annu. Rev. Genet. 2018, 52, 43–63. [Google Scholar] [CrossRef]
  179. Teddy, J.M.; Kulesa, P.M. In vivo evidence for short- and long-range cell communication in cranial neural crest cells. Development 2004, 131, 6141–6151. [Google Scholar] [CrossRef]
  180. Huang, G.Y.; Cooper, E.S.; Waldo, K.; Kirby, M.L.; Gilula, N.B.; Lo, C.W. Gap junction-mediated cell-cell communication modulates mouse neural crest migration. J. Cell Biol. 1998, 143, 1725–1734. [Google Scholar] [CrossRef]
  181. Sullivan, R.; Huang, G.Y.; Meyer, R.A.; Wessels, A.; Linask, K.K.; Lo, C.W. Heart malformations in transgenic mice exhibiting dominant negative inhibition of gap junctional communication in neural crest cells. Dev. Biol. 1998, 204, 224–234. [Google Scholar] [CrossRef] [PubMed]
  182. Lo, C.W.; Waldo, K.L.; Kirby, M.L. Gap junction communication and the modulation of cardiac neural crest cells. Trends Cardiovasc. Med. 1999, 9, 63–69. [Google Scholar] [CrossRef] [PubMed]
  183. Lo, C.W.; Cohen, M.F.; Huang, G.Y.; Lazatin, B.O.; Patel, N.; Sullivan, R.; Pauken, C.; Park, S.M. Cx43 gap junction gene expression and gap junctional communication in mouse neural crest cells. Dev. Genet. 1997, 20, 119–132. [Google Scholar] [CrossRef]
  184. Gustafson, C.M.; Roffers-Agarwal, J.; Gammill, L.S. Chick cranial neural crest cells release extracellular vesicles that are critical for their migration. J. Cell Sci. 2022, 135, jcs260272. [Google Scholar] [CrossRef]
  185. Nakagawa, S.; Takeichi, M. Neural crest emigration from the neural tube depends on regulated cadherin expression. Development 1998, 125, 2963–2971. [Google Scholar] [CrossRef]
  186. Xu, X.; Li, W.E.; Huang, G.Y.; Meyer, R.; Chen, T.; Luo, Y.; Thomas, M.P.; Radice, G.L.; Lo, C.W. Modulation of mouse neural crest cell motility by N-cadherin and connexin 43 gap junctions. J. Cell Biol. 2001, 154, 217–230. [Google Scholar] [CrossRef]
  187. Fishwick, K.J.; Neiderer, T.E.; Jhingory, S.; Bronner, M.E.; Taneyhill, L.A. The tight junction protein claudin-1 influences cranial neural crest cell emigration. Mech. Dev. 2012, 129, 275–283. [Google Scholar] [CrossRef]
  188. Wu, C.Y.; Jhingory, S.; Taneyhill, L.A. The tight junction scaffolding protein cingulin regulates neural crest cell migration. Dev. Dyn. 2011, 240, 2309–2323. [Google Scholar] [CrossRef]
  189. Vacca, B.; Sanchez-Heras, E.; Steed, E.; Busson, S.L.; Balda, M.S.; Ohnuma, S.I.; Sasai, N.; Mayor, R.; Matter, K. Control of neural crest induction by MarvelD3-mediated attenuation of JNK signalling. Sci. Rep. 2018, 8, 1204. [Google Scholar] [CrossRef]
  190. Smith, A.; Robinson, V.; Patel, K.; Wilkinson, D.G. The EphA4 and EphB1 receptor tyrosine kinases and ephrin-B2 ligand regulate targeted migration of branchial neural crest cells. Curr. Biol. 1997, 7, 561–570. [Google Scholar] [CrossRef]
  191. Bannerman, P.; Nichols, W.; Puhalla, S.; Oliver, T.; Berman, M.; Pleasure, D. Early migratory rat neural crest cells express functional gap junctions: Evidence that neural crest cell survival requires gap junction function. J. Neurosci. Res. 2000, 61, 605–615. [Google Scholar] [CrossRef] [PubMed]
  192. Bhattacharya, D.; Khan, B.; Simoes-Costa, M. Neural crest metabolism: At the crossroads of development and disease. Dev. Biol. 2021, 475, 245–255. [Google Scholar] [CrossRef]
  193. Wang, G.; Chen, E.N.; Liang, C.; Liang, J.; Gao, L.R.; Chuai, M.; Munsterberg, A.; Bao, Y.; Cao, L.; Yang, X. Atg7-Mediated Autophagy Is Involved in the Neural Crest Cell Generation in Chick Embryo. Mol. Neurobiol. 2018, 55, 3523–3536. [Google Scholar] [CrossRef] [PubMed]
  194. Wang, X.Y.; Li, S.; Wang, G.; Ma, Z.L.; Chuai, M.; Cao, L.; Yang, X. High glucose environment inhibits cranial neural crest survival by activating excessive autophagy in the chick embryo. Sci. Rep. 2015, 5, 18321. [Google Scholar] [CrossRef] [PubMed]
  195. Clevers, H. The intestinal crypt, a prototype stem cell compartment. Cell 2013, 154, 274–284. [Google Scholar] [CrossRef] [PubMed]
  196. Clevers, H.; Batlle, E. SnapShot: The intestinal crypt. Cell 2013, 152, 1198–1198.e2. [Google Scholar] [CrossRef] [PubMed]
  197. Maloy, K.J.; Powrie, F. Intestinal homeostasis and its breakdown in inflammatory bowel disease. Nature 2011, 474, 298–306. [Google Scholar] [CrossRef] [PubMed]
  198. Eisenhoffer, G.T.; Loftus, P.D.; Yoshigi, M.; Otsuna, H.; Chien, C.B.; Morcos, P.A.; Rosenblatt, J. Crowding induces live cell extrusion to maintain homeostatic cell numbers in epithelia. Nature 2012, 484, 546–549. [Google Scholar] [CrossRef]
  199. Patterson, A.M.; Watson, A.J.M. Deciphering the Complex Signaling Systems That Regulate Intestinal Epithelial Cell Death Processes and Shedding. Front. Immunol. 2017, 8, 841. [Google Scholar] [CrossRef]
  200. Ngo, P.A.; Neurath, M.F.; Lopez-Posadas, R. Impact of Epithelial Cell Shedding on Intestinal Homeostasis. Int. J. Mol. Sci. 2022, 23, 4160. [Google Scholar] [CrossRef]
  201. Rosenblatt, J.; Raff, M.C.; Cramer, L.P. An epithelial cell destined for apoptosis signals its neighbors to extrude it by an actin- and myosin-dependent mechanism. Curr. Biol. 2001, 11, 1847–1857. [Google Scholar] [CrossRef]
  202. Marchiando, A.M.; Shen, L.; Graham, W.V.; Edelblum, K.L.; Duckworth, C.A.; Guan, Y.; Montrose, M.H.; Turner, J.R.; Watson, A.J. The epithelial barrier is maintained by in vivo tight junction expansion during pathologic intestinal epithelial shedding. Gastroenterology 2011, 140, 1208–1218. [Google Scholar] [CrossRef] [PubMed]
  203. Zihni, C.; Mills, C.; Matter, K.; Balda, M.S. Tight junctions: From simple barriers to multifunctional molecular gates. Nat. Rev. Mol. Cell Biol. 2016, 17, 564–580. [Google Scholar] [CrossRef] [PubMed]
  204. Martin, T.A. The role of tight junctions in cancer metastasis. Semin Cell Dev. Biol. 2014, 36, 224–231. [Google Scholar] [CrossRef] [PubMed]
  205. Martin, T.A.; Jiang, W.G. Loss of tight junction barrier function and its role in cancer metastasis. Biochim. Biophys. Acta 2009, 1788, 872–891. [Google Scholar] [CrossRef] [PubMed]
  206. Nehme, Z.; Roehlen, N.; Dhawan, P.; Baumert, T.F. Tight Junction Protein Signaling and Cancer Biology. Cells 2023, 12, 243. [Google Scholar] [CrossRef] [PubMed]
  207. Kyuno, D.; Takasawa, A.; Kikuchi, S.; Takemasa, I.; Osanai, M.; Kojima, T. Role of tight junctions in the epithelial-to-mesenchymal transition of cancer cells. Biochim. Biophys. Acta Biomembr. 2021, 1863, 183503. [Google Scholar] [CrossRef] [PubMed]
  208. Aijaz, S.; Balda, M.S.; Matter, K. Tight junctions: Molecular architecture and function. Int. Rev. Cytol. 2006, 248, 261–298. [Google Scholar] [CrossRef]
  209. Cummins, P.M. Occludin: One protein, many forms. Mol. Cell Biol. 2012, 32, 242–250. [Google Scholar] [CrossRef]
  210. Steed, E.; Balda, M.S.; Matter, K. Dynamics and functions of tight junctions. Trends Cell Biol. 2010, 20, 142–149. [Google Scholar] [CrossRef]
  211. Bhat, A.A.; Uppada, S.; Achkar, I.W.; Hashem, S.; Yadav, S.K.; Shanmugakonar, M.; Al-Naemi, H.A.; Haris, M.; Uddin, S. Tight Junction Proteins and Signaling Pathways in Cancer and Inflammation: A Functional Crosstalk. Front. Physiol. 2018, 9, 1942. [Google Scholar] [CrossRef] [PubMed]
  212. Garcia, M.A.; Nelson, W.J.; Chavez, N. Cell-Cell Junctions Organize Structural and Signaling Networks. Cold Spring Harb. Perspect. Biol. 2018, 10, a029181. [Google Scholar] [CrossRef] [PubMed]
  213. Singh, A.B.; Uppada, S.B.; Dhawan, P. Claudin proteins, outside-in signaling, and carcinogenesis. Pflug. Arch. 2017, 469, 69–75. [Google Scholar] [CrossRef] [PubMed]
  214. Franke, W.W. Discovering the molecular components of intercellular junctions—A historical view. Cold Spring Harb. Perspect. Biol. 2009, 1, a003061. [Google Scholar] [CrossRef] [PubMed]
  215. Rubsam, M.; Broussard, J.A.; Wickstrom, S.A.; Nekrasova, O.; Green, K.J.; Niessen, C.M. Adherens Junctions and Desmosomes Coordinate Mechanics and Signaling to Orchestrate Tissue Morphogenesis and Function: An Evolutionary Perspective. Cold Spring Harb. Perspect. Biol. 2018, 10, a029207. [Google Scholar] [CrossRef] [PubMed]
  216. Green, K.J.; Roth-Carter, Q.; Niessen, C.M.; Nichols, S.A. Tracing the Evolutionary Origin of Desmosomes. Curr. Biol. 2020, 30, R535–R543. [Google Scholar] [CrossRef] [PubMed]
  217. Madara, J.L. Maintenance of the macromolecular barrier at cell extrusion sites in intestinal epithelium: Physiological rearrangement of tight junctions. J. Membr. Biol. 1990, 116, 177–184. [Google Scholar] [CrossRef] [PubMed]
  218. Turner, J.R. Intestinal mucosal barrier function in health and disease. Nat. Rev. Immunol. 2009, 9, 799–809. [Google Scholar] [CrossRef]
  219. Cliffe, L.J.; Humphreys, N.E.; Lane, T.E.; Potten, C.S.; Booth, C.; Grencis, R.K. Accelerated intestinal epithelial cell turnover: A new mechanism of parasite expulsion. Science 2005, 308, 1463–1465. [Google Scholar] [CrossRef]
  220. Nenci, A.; Becker, C.; Wullaert, A.; Gareus, R.; van Loo, G.; Danese, S.; Huth, M.; Nikolaev, A.; Neufert, C.; Madison, B.; et al. Epithelial NEMO links innate immunity to chronic intestinal inflammation. Nature 2007, 446, 557–561. [Google Scholar] [CrossRef]
  221. Clayburgh, D.R.; Shen, L.; Turner, J.R. A porous defense: The leaky epithelial barrier in intestinal disease. Lab. Investig. 2004, 84, 282–291. [Google Scholar] [CrossRef]
  222. Bruewer, M.; Luegering, A.; Kucharzik, T.; Parkos, C.A.; Madara, J.L.; Hopkins, A.M.; Nusrat, A. Proinflammatory cytokines disrupt epithelial barrier function by apoptosis-independent mechanisms. J. Immunol. 2003, 171, 6164–6172. [Google Scholar] [CrossRef]
  223. Gunther, C.; Neumann, H.; Neurath, M.F.; Becker, C. Apoptosis, necrosis and necroptosis: Cell death regulation in the intestinal epithelium. Gut 2013, 62, 1062–1071. [Google Scholar] [CrossRef] [PubMed]
  224. Patankar, J.V.; Becker, C. Cell death in the gut epithelium and implications for chronic inflammation. Nat. Rev. Gastroenterol. Hepatol. 2020, 17, 543–556. [Google Scholar] [CrossRef]
  225. Ayyar, K.K.; Moss, A.C. Exosomes in Intestinal Inflammation. Front. Pharmacol. 2021, 12, 658505. [Google Scholar] [CrossRef]
  226. van Niel, G.; Raposo, G.; Candalh, C.; Boussac, M.; Hershberg, R.; Cerf-Bensussan, N.; Heyman, M. Intestinal epithelial cells secrete exosome-like vesicles. Gastroenterology 2001, 121, 337–349. [Google Scholar] [CrossRef] [PubMed]
  227. Mallegol, J.; Van Niel, G.; Lebreton, C.; Lepelletier, Y.; Candalh, C.; Dugave, C.; Heath, J.K.; Raposo, G.; Cerf-Bensussan, N.; Heyman, M. T84-intestinal epithelial exosomes bear MHC class II/peptide complexes potentiating antigen presentation by dendritic cells. Gastroenterology 2007, 132, 1866–1876. [Google Scholar] [CrossRef] [PubMed]
  228. Canbay, A.; Feldstein, A.E.; Higuchi, H.; Werneburg, N.; Grambihler, A.; Bronk, S.F.; Gores, G.J. Kupffer cell engulfment of apoptotic bodies stimulates death ligand and cytokine expression. Hepatology 2003, 38, 1188–1198. [Google Scholar] [CrossRef] [PubMed]
  229. Xie, Y.; Zhao, Y.; Shi, L.; Li, W.; Chen, K.; Li, M.; Chen, X.; Zhang, H.; Li, T.; Matsuzawa-Ishimoto, Y.; et al. Gut epithelial TSC1/mTOR controls RIPK3-dependent necroptosis in intestinal inflammation and cancer. J. Clin. Investig. 2020, 130, 2111–2128. [Google Scholar] [CrossRef] [PubMed]
  230. Xu, J.; Li, S.; Jin, W.; Zhou, H.; Zhong, T.; Cheng, X.; Fu, Y.; Xiao, P.; Cheng, H.; Wang, D.; et al. Epithelial Gab1 calibrates RIPK3-dependent necroptosis to prevent intestinal inflammation. JCI Insight 2023, 8, e162701. [Google Scholar] [CrossRef] [PubMed]
  231. Churchill, M.J.; Mitchell, P.S.; Rauch, I. Epithelial Pyroptosis in Host Defense. J. Mol. Biol. 2022, 434, 167278. [Google Scholar] [CrossRef]
  232. Rauch, I.; Deets, K.A.; Ji, D.X.; von Moltke, J.; Tenthorey, J.L.; Lee, A.Y.; Philip, N.H.; Ayres, J.S.; Brodsky, I.E.; Gronert, K.; et al. NAIP-NLRC4 Inflammasomes Coordinate Intestinal Epithelial Cell Expulsion with Eicosanoid and IL-18 Release via Activation of Caspase-1 and -8. Immunity 2017, 46, 649–659. [Google Scholar] [CrossRef] [PubMed]
  233. He, W.T.; Wan, H.; Hu, L.; Chen, P.; Wang, X.; Huang, Z.; Yang, Z.H.; Zhong, C.Q.; Han, J. Gasdermin D is an executor of pyroptosis and required for interleukin-1beta secretion. Cell Res. 2015, 25, 1285–1298. [Google Scholar] [CrossRef] [PubMed]
  234. Lin, P.H.; Lin, H.Y.; Kuo, C.C.; Yang, L.T. N-terminal functional domain of Gasdermin A3 regulates mitochondrial homeostasis via mitochondrial targeting. J. Biomed. Sci. 2015, 22, 44. [Google Scholar] [CrossRef]
  235. Shi, P.; Tang, A.; Xian, L.; Hou, S.; Zou, D.; Lv, Y.; Huang, Z.; Wang, Q.; Song, A.; Lin, Z.; et al. Loss of conserved Gsdma3 self-regulation causes autophagy and cell death. Biochem. J. 2015, 468, 325–336. [Google Scholar] [CrossRef] [PubMed]
  236. Tamura, M.; Shiroishi, T. GSDM family genes meet autophagy. Biochem. J. 2015, 469, e5–e7. [Google Scholar] [CrossRef]
  237. Privitera, G.; Rana, N.; Armuzzi, A.; Pizarro, T.T. The gasdermin protein family: Emerging roles in gastrointestinal health and disease. Nat. Rev. Gastroenterol. Hepatol. 2023, 20, 366–387. [Google Scholar] [CrossRef] [PubMed]
  238. Randall-Demllo, S.; Chieppa, M.; Eri, R. Intestinal epithelium and autophagy: Partners in gut homeostasis. Front. Immunol. 2013, 4, 301. [Google Scholar] [CrossRef]
  239. Benjamin, J.L.; Sumpter, R., Jr.; Levine, B.; Hooper, L.V. Intestinal epithelial autophagy is essential for host defense against invasive bacteria. Cell Host Microbe 2013, 13, 723–734. [Google Scholar] [CrossRef]
  240. Telpaz, S.; Bel, S. Autophagy in intestinal epithelial cells prevents gut inflammation. Trends Cell Biol. 2023, 33, 817–819. [Google Scholar] [CrossRef]
  241. Chen, S.L.; Li, C.M.; Li, W.; Liu, Q.S.; Hu, S.Y.; Zhao, M.Y.; Hu, D.S.; Hao, Y.W.; Zeng, J.H.; Zhang, Y. How autophagy, a potential therapeutic target, regulates intestinal inflammation. Front. Immunol. 2023, 14, 1087677. [Google Scholar] [CrossRef] [PubMed]
  242. 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]
  243. Solanas, G.; Batlle, E. Control of cell adhesion and compartmentalization in the intestinal epithelium. Exp. Cell Res. 2011, 317, 2695–2701. [Google Scholar] [CrossRef] [PubMed]
  244. Holmberg, J.; Genander, M.; Halford, M.M.; Anneren, C.; Sondell, M.; Chumley, M.J.; Silvany, R.E.; Henkemeyer, M.; Frisen, J. EphB receptors coordinate migration and proliferation in the intestinal stem cell niche. Cell 2006, 125, 1151–1163. [Google Scholar] [CrossRef] [PubMed]
  245. Anderton, M.; van der Meulen, E.; Blumenthal, M.J.; Schafer, G. The Role of the Eph Receptor Family in Tumorigenesis. Cancers 2021, 13, 206. [Google Scholar] [CrossRef] [PubMed]
  246. Janes, P.W.; Vail, M.E.; Ernst, M.; Scott, A.M. Eph Receptors in the Immunosuppressive Tumor Microenvironment. Cancer Res. 2021, 81, 801–805. [Google Scholar] [CrossRef] [PubMed]
  247. Zhou, X.; Tu, P.; Chen, X.; Guo, S.; Wang, J. Eph Receptors: Actors in Tumor Microenvironment. Crit. Rev. Oncog. 2017, 22, 499–505. [Google Scholar] [CrossRef]
  248. Cowan, C.A.; Henkemeyer, M. Ephrins in reverse, park and drive. Trends Cell. Biol. 2002, 12, 339–346. [Google Scholar] [CrossRef]
  249. Merlos-Suarez, A.; Batlle, E. Eph-ephrin signalling in adult tissues and cancer. Curr. Opin. Cell Biol. 2008, 20, 194–200. [Google Scholar] [CrossRef]
  250. Genander, M.; Frisen, J. Eph receptors tangled up in two: Independent control of cell positioning and proliferation. Cell Cycle 2010, 9, 1865–1866. [Google Scholar] [CrossRef]
  251. Amessou, M.; Kandouz, M. Role of the Family of Ephs and Ephrins in Cell-Cell Communication in Cancer. In Intercellular Communication in Cancer; Kandouz, M., Ed.; Springer: Dordrecht, The Netherlands, 2015; pp. 255–286. [Google Scholar]
  252. Kandouz, M. The Eph/Ephrin family in cancer metastasis: Communication at the service of invasion. Cancer Metastasis Rev. 2012, 31, 353–373. [Google Scholar] [CrossRef]
  253. Pasquale, E.B. Eph receptors and ephrins in cancer progression. Nat. Rev. Cancer 2024, 24, 5–27. [Google Scholar] [CrossRef] [PubMed]
  254. Batlle, E.; Bacani, J.; Begthel, H.; Jonkheer, S.; Gregorieff, A.; van de Born, M.; Malats, N.; Sancho, E.; Boon, E.; Pawson, T.; et al. EphB receptor activity suppresses colorectal cancer progression. Nature 2005, 435, 1126–1130. [Google Scholar] [CrossRef]
  255. Batlle, E.; Henderson, J.T.; Beghtel, H.; van den Born, M.M.; Sancho, E.; Huls, G.; Meeldijk, J.; Robertson, J.; van de Wetering, M.; Pawson, T.; et al. Beta-catenin and TCF mediate cell positioning in the intestinal epithelium by controlling the expression of EphB/ephrinB. Cell 2002, 111, 251–263. [Google Scholar] [CrossRef] [PubMed]
  256. van de Wetering, M.; Sancho, E.; Verweij, C.; de Lau, W.; Oving, I.; Hurlstone, A.; van der Horn, K.; Batlle, E.; Coudreuse, D.; Haramis, A.P.; et al. The beta-catenin/TCF-4 complex imposes a crypt progenitor phenotype on colorectal cancer cells. Cell 2002, 111, 241–250. [Google Scholar] [CrossRef] [PubMed]
  257. Moser, A.R.; Pitot, H.C.; Dove, W.F. A dominant mutation that predisposes to multiple intestinal neoplasia in the mouse. Science 1990, 247, 322–324. [Google Scholar] [CrossRef] [PubMed]
  258. Jubb, A.M.; Zhong, F.; Bheddah, S.; Grabsch, H.I.; Frantz, G.D.; Mueller, W.; Kavi, V.; Quirke, P.; Polakis, P.; Koeppen, H. EphB2 is a prognostic factor in colorectal cancer. Clin. Cancer Res. 2005, 11, 5181–5187. [Google Scholar] [CrossRef] [PubMed]
  259. Guo, D.L.; Zhang, J.; Yuen, S.T.; Tsui, W.Y.; Chan, A.S.; Ho, C.; Ji, J.; Leung, S.Y.; Chen, X. Reduced expression of EphB2 that parallels invasion and metastasis in colorectal tumours. Carcinogenesis 2006, 27, 454–464. [Google Scholar] [CrossRef] [PubMed]
  260. Lugli, A.; Spichtin, H.; Maurer, R.; Mirlacher, M.; Kiefer, J.; Huusko, P.; Azorsa, D.; Terracciano, L.; Sauter, G.; Kallioniemi, O.P.; et al. EphB2 expression across 138 human tumor types in a tissue microarray: High levels of expression in gastrointestinal cancers. Clin. Cancer Res. 2005, 11, 6450–6458. [Google Scholar] [CrossRef]
  261. Cortina, C.; Palomo-Ponce, S.; Iglesias, M.; Fernandez-Masip, J.L.; Vivancos, A.; Whissell, G.; Huma, M.; Peiro, N.; Gallego, L.; Jonkheer, S.; et al. EphB-ephrin-B interactions suppress colorectal cancer progression by compartmentalizing tumor cells. Nat. Genet. 2007, 39, 1376–1383. [Google Scholar] [CrossRef]
  262. Chiu, S.T.; Chang, K.J.; Ting, C.H.; Shen, H.C.; Li, H.; Hsieh, F.J. Over-expression of EphB3 enhances cell-cell contacts and suppresses tumor growth in HT-29 human colon cancer cells. Carcinogenesis 2009, 30, 1475–1486. [Google Scholar] [CrossRef]
  263. 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]
  264. Chukkapalli, S.; Amessou, M.; Dilly, A.K.; Dekhil, H.; Zhao, J.; Liu, Q.; Bejna, A.; Thomas, R.D.; Bandyopadhyay, S.; Bismar, T.A.; et al. Role of the EphB2 receptor in autophagy, apoptosis and invasion in human breast cancer cells. Exp. Cell Res. 2014, 320, 233–246. [Google Scholar] [CrossRef] [PubMed]
  265. Kandouz, M.; Haidara, K.; Zhao, J.; Brisson, M.L.; Batist, G. The EphB2 tumor suppressor induces autophagic cell death via concomitant activation of the ERK1/2 and PI3K pathways. Cell Cycle 2010, 9, 398–407. [Google Scholar] [CrossRef] [PubMed]
  266. Zhang, H.; Cui, Z.; Cheng, D.; Du, Y.; Guo, X.; Gao, R.; Chen, J.; Sun, W.; He, R.; Ma, X.; et al. RNF186 regulates EFNB1 (ephrin B1)-EPHB2-induced autophagy in the colonic epithelial cells for the maintenance of intestinal homeostasis. Autophagy 2021, 17, 3030–3047. [Google Scholar] [CrossRef] [PubMed]
  267. Tanabe, H.; Kuribayashi, K.; Tsuji, N.; Tanaka, M.; Kobayashi, D.; Watanabe, N. Sesamin induces autophagy in colon cancer cells by reducing tyrosine phosphorylation of EphA1 and EphB2. Int. J. Oncol. 2011, 39, 33–40. [Google Scholar] [CrossRef] [PubMed]
  268. Qiu, P.; Li, D.; Xiao, C.; Xu, F.; Chen, X.; Chang, Y.; Liu, L.; Zhang, L.; Zhao, Q.; Chen, Y. The Eph/ephrin system symphony of gut inflammation. Pharmacol. Res. 2023, 197, 106976. [Google Scholar] [CrossRef]
  269. Darling, T.K.; Lamb, T.J. Emerging Roles for Eph Receptors and Ephrin Ligands in Immunity. Front. Immunol. 2019, 10, 1473. [Google Scholar] [CrossRef] [PubMed]
  270. Kuijper, S.; Turner, C.J.; Adams, R.H. Regulation of angiogenesis by Eph-ephrin interactions. Trends Cardiovasc. Med. 2007, 17, 145–151. [Google Scholar] [CrossRef] [PubMed]
  271. Shiuan, E.; Chen, J. Eph Receptor Tyrosine Kinases in Tumor Immunity. Cancer Res. 2016, 76, 6452–6457. [Google Scholar] [CrossRef]
  272. El Khoury, F.; Corcos, L.; Durand, S.; Simon, B.; Le Jossic-Corcos, C. Acquisition of anticancer drug resistance is partially associated with cancer stemness in human colon cancer cells. Int. J. Oncol. 2016, 49, 2558–2568. [Google Scholar] [CrossRef]
  273. Park, S.H.; Jo, M.J.; Kim, B.R.; Jeong, Y.A.; Na, Y.J.; Kim, J.L.; Jeong, S.; Yun, H.K.; Kim, D.Y.; Kim, B.G.; et al. Sonic hedgehog pathway activation is associated with cetuximab resistance and EPHB3 receptor induction in colorectal cancer. Theranostics 2019, 9, 2235–2251. [Google Scholar] [CrossRef]
  274. Primeaux, M.; Liu, X.; Gowrikumar, S.; Fatima, I.; Fisher, K.W.; Bastola, D.; Vecchio, A.J.; Singh, A.B.; Dhawan, P. Claudin-1 interacts with EPHA2 to promote cancer stemness and chemoresistance in colorectal cancer. Cancer Lett. 2023, 579, 216479. [Google Scholar] [CrossRef] [PubMed]
  275. DiPrima, M.; Wang, D.; Troster, A.; Maric, D.; Terrades-Garcia, N.; Ha, T.; Kwak, H.; Sanchez-Martin, D.; Kudlinzki, D.; Schwalbe, H.; et al. Identification of Eph receptor signaling as a regulator of autophagy and a therapeutic target in colorectal carcinoma. Mol. Oncol. 2019, 13, 2441–2459. [Google Scholar] [CrossRef] [PubMed]
  276. Troster, A.; DiPrima, M.; Jores, N.; Kudlinzki, D.; Sreeramulu, S.; Gande, S.L.; Linhard, V.; Ludig, D.; Schug, A.; Saxena, K.; et al. Optimization of the Lead Compound NVP-BHG712 as a Colorectal Cancer Inhibitor. Chemistry 2023, 29, e202203967. [Google Scholar] [CrossRef] [PubMed]
  277. Troster, A.; Jores, N.; Mineev, K.S.; Sreeramulu, S.; DiPrima, M.; Tosato, G.; Schwalbe, H. Targeting EPHA2 with Kinase Inhibitors in Colorectal Cancer. ChemMedChem 2023, 18, e202300420. [Google Scholar] [CrossRef]
  278. Alam, S.K.; Yadav, V.K.; Bajaj, S.; Datta, A.; Dutta, S.K.; Bhattacharyya, M.; Bhattacharya, S.; Debnath, S.; Roy, S.; Boardman, L.A.; et al. DNA damage-induced ephrin-B2 reverse signaling promotes chemoresistance and drives EMT in colorectal carcinoma harboring mutant p53. Cell Death Differ. 2016, 23, 707–722. [Google Scholar] [CrossRef] [PubMed]
  279. Cioce, M.; Fazio, V.M. EphA2 and EGFR: Friends in Life, Partners in Crime. Can EphA2 Be a Predictive Biomarker of Response to Anti-EGFR Agents? Cancers 2021, 13, 700. [Google Scholar] [CrossRef]
  280. Strimpakos, A.; Pentheroudakis, G.; Kotoula, V.; De Roock, W.; Kouvatseas, G.; Papakostas, P.; Makatsoris, T.; Papamichael, D.; Andreadou, A.; Sgouros, J.; et al. The prognostic role of ephrin A2 and endothelial growth factor receptor pathway mediators in patients with advanced colorectal cancer treated with cetuximab. Clin. Colorectal. Cancer 2013, 12, 267–274. [Google Scholar] [CrossRef] [PubMed]
  281. Cuyas, E.; Queralt, B.; Martin-Castillo, B.; Bosch-Barrera, J.; Menendez, J.A. EphA2 receptor activation with ephrin-A1 ligand restores cetuximab efficacy in NRAS-mutant colorectal cancer cells. Oncol. Rep. 2017, 38, 263–270. [Google Scholar] [CrossRef]
  282. Lau, A.; Le, N.; Nguyen, C.; Kandpal, R.P. Signals transduced by Eph receptors and ephrin ligands converge on MAP kinase and AKT pathways in human cancers. Cell. Signal. 2023, 104, 110579. [Google Scholar] [CrossRef] [PubMed]
  283. Martini, G.; Cardone, C.; Vitiello, P.P.; Belli, V.; Napolitano, S.; Troiani, T.; Ciardiello, D.; Della Corte, C.M.; Morgillo, F.; Matrone, N.; et al. EPHA2 Is a Predictive Biomarker of Resistance and a Potential Therapeutic Target for Improving Antiepidermal Growth Factor Receptor Therapy in Colorectal Cancer. Mol. Cancer Ther. 2019, 18, 845–855. [Google Scholar] [CrossRef] [PubMed]
  284. Bhatia, S.; Sharma, J.; Bukkapatnam, S.; Oweida, A.; Lennon, S.; Phan, A.; Milner, D.; Uyanga, N.; Jimeno, A.; Raben, D.; et al. Inhibition of EphB4-Ephrin-B2 Signaling Enhances Response to Cetuximab-Radiation Therapy in Head and Neck Cancers. Clin. Cancer Res. 2018, 24, 4539–4550. [Google Scholar] [CrossRef] [PubMed]
  285. Stahl, S.; Kaminskyy, V.O.; Efazat, G.; Hyrslova Vaculova, A.; Rodriguez-Nieto, S.; Moshfegh, A.; Lewensohn, R.; Viktorsson, K.; Zhivotovsky, B. Inhibition of Ephrin B3-mediated survival signaling contributes to increased cell death response of non-small cell lung carcinoma cells after combined treatment with ionizing radiation and PKC 412. Cell Death Dis. 2013, 4, e454. [Google Scholar] [CrossRef] [PubMed]
  286. Beekhof, R.; Bertotti, A.; Bottger, F.; Vurchio, V.; Cottino, F.; Zanella, E.R.; Migliardi, G.; Viviani, M.; Grassi, E.; Lupo, B.; et al. Phosphoproteomics of patient-derived xenografts identifies targets and markers associated with sensitivity and resistance to EGFR blockade in colorectal cancer. Sci. Transl. Med. 2023, 15, eabm3687. [Google Scholar] [CrossRef]
  287. Yang, S.Y.; Sales, K.M.; Fuller, B.; Seifalian, A.M.; Winslet, M.C. Apoptosis and colorectal cancer: Implications for therapy. Trends Mol. Med. 2009, 15, 225–233. [Google Scholar] [CrossRef] [PubMed]
  288. Watson, A.J. Apoptosis and colorectal cancer. Gut 2004, 53, 1701–1709. [Google Scholar] [CrossRef] [PubMed]
  289. Sinicrope, F.A.; Roddey, G.; McDonnell, T.J.; Shen, Y.; Cleary, K.R.; Stephens, L.C. Increased apoptosis accompanies neoplastic development in the human colorectum. Clin. Cancer Res. 1996, 2, 1999–2006. [Google Scholar]
  290. Manzoor, S.; Muhammad, J.S.; Maghazachi, A.A.; Hamid, Q. Autophagy: A Versatile Player in the Progression of Colorectal Cancer and Drug Resistance. Front. Oncol. 2022, 12, 924290. [Google Scholar] [CrossRef]
  291. Xie, Q.; Liu, Y.; Li, X. The interaction mechanism between autophagy and apoptosis in colon cancer. Transl. Oncol. 2020, 13, 100871. [Google Scholar] [CrossRef]
  292. de la Cruz-Morcillo, M.A.; Valero, M.L.; Callejas-Valera, J.L.; Arias-Gonzalez, L.; Melgar-Rojas, P.; Galan-Moya, E.M.; Garcia-Gil, E.; Garcia-Cano, J.; Sanchez-Prieto, R. P38MAPK is a major determinant of the balance between apoptosis and autophagy triggered by 5-fluorouracil: Implication in resistance. Oncogene 2012, 31, 1073–1085. [Google Scholar] [CrossRef]
  293. Chen, Z.; Gao, S.; Wang, D.; Song, D.; Feng, Y. Colorectal cancer cells are resistant to anti-EGFR monoclonal antibody through adapted autophagy. Am. J. Transl. Res. 2016, 8, 1190–1196. [Google Scholar] [PubMed]
  294. Koustas, E.; Karamouzis, M.V.; Mihailidou, C.; Schizas, D.; Papavassiliou, A.G. Co-targeting of EGFR and autophagy signaling is an emerging treatment strategy in metastatic colorectal cancer. Cancer Lett. 2017, 396, 94–102. [Google Scholar] [CrossRef]
  295. Guo, Y.; Shi, W.; Fang, R. miR-18a-5p promotes melanoma cell proliferation and inhibits apoptosis and autophagy by targeting EPHA7 signaling. Mol. Med. Rep. 2021, 23. [Google Scholar] [CrossRef] [PubMed]
  296. Yan, H.; Talty, R.; Aladelokun, O.; Bosenberg, M.; Johnson, C.H. Ferroptosis in colorectal cancer: A future target? Br. J. Cancer 2023, 128, 1439–1451. [Google Scholar] [CrossRef] [PubMed]
  297. Liang, H.; He, X.; Tong, Y.; Bai, N.; Pu, Y.; Han, K.; Wang, Y. Ferroptosis open a new door for colorectal cancer treatment. Front. Oncol. 2023, 13, 1059520. [Google Scholar] [CrossRef] [PubMed]
  298. Wang, Y.; Zhang, Z.; Sun, W.; Zhang, J.; Xu, Q.; Zhou, X.; Mao, L. Ferroptosis in colorectal cancer: Potential mechanisms and effective therapeutic targets. Biomed. Pharmacother. 2022, 153, 113524. [Google Scholar] [CrossRef]
  299. Chaudhary, N.; Choudhary, B.S.; Shah, S.G.; Khapare, N.; Dwivedi, N.; Gaikwad, A.; Joshi, N.; Raichanna, J.; Basu, S.; Gurjar, M.; et al. Lipocalin 2 expression promotes tumor progression and therapy resistance by inhibiting ferroptosis in colorectal cancer. Int. J. Cancer 2021, 149, 1495–1511. [Google Scholar] [CrossRef]
  300. Lin, Y.P.; Wu, J.I.; Tseng, C.W.; Chen, H.J.; Wang, L.H. Gjb4 serves as a novel biomarker for lung cancer and promotes metastasis and chemoresistance via Src activation. Oncogene 2019, 38, 822–837. [Google Scholar] [CrossRef]
  301. Yang, J.; Qin, G.; Luo, M.; Chen, J.; Zhang, Q.; Li, L.; Pan, L.; Qin, S. Reciprocal positive regulation between Cx26 and PI3K/Akt pathway confers acquired gefitinib resistance in NSCLC cells via GJIC-independent induction of EMT. Cell Death Dis. 2015, 6, e1829. [Google Scholar] [CrossRef]
  302. Gielen, P.R.; Aftab, Q.; Ma, N.; Chen, V.C.; Hong, X.; Lozinsky, S.; Naus, C.C.; Sin, W.C. Connexin43 confers Temozolomide resistance in human glioma cells by modulating the mitochondrial apoptosis pathway. Neuropharmacology 2013, 75, 539–548. [Google Scholar] [CrossRef]
  303. Ni, C.; Lou, X.; Yao, X.; Wang, L.; Wan, J.; Duan, X.; Liang, J.; Zhang, K.; Yang, Y.; Zhang, L.; et al. ZIP1(+) fibroblasts protect lung cancer against chemotherapy via connexin-43 mediated intercellular Zn(2+) transfer. Nat. Commun. 2022, 13, 5919. [Google Scholar] [CrossRef]
  304. Arora, S.; Heyza, J.R.; Chalfin, E.C.; Ruch, R.J.; Patrick, S.M. Gap Junction Intercellular Communication Positively Regulates Cisplatin Toxicity by Inducing DNA Damage through Bystander Signaling. Cancers 2018, 10, 368. [Google Scholar] [CrossRef]
  305. Blondy, S.; David, V.; Verdier, M.; Mathonnet, M.; Perraud, A.; Christou, N. 5-Fluorouracil resistance mechanisms in colorectal cancer: From classical pathways to promising processes. Cancer Sci. 2020, 111, 3142–3154. [Google Scholar] [CrossRef] [PubMed]
  306. Han, Y.; Wang, H.; Chen, H.; Tan, T.; Wang, Y.; Yang, H.; Ding, Y.; Wang, S. CX43 down-regulation promotes cell aggressiveness and 5-fluorouracil-resistance by attenuating cell stiffness in colorectal carcinoma. Cancer Biol. Ther. 2023, 24, 2221879. [Google Scholar] [CrossRef] [PubMed]
  307. Zou, Z.W.; Chen, H.J.; Yu, J.L.; Huang, Z.H.; Fang, S.; Lin, X.H. Gap junction composed of connexin43 modulates 5-fluorouracil, oxaliplatin and irinotecan resistance on colorectal cancers. Mol. Med. Rep. 2016, 14, 4893–4900. [Google Scholar] [CrossRef] [PubMed]
  308. Buhrmann, C.; Shayan, P.; Kraehe, P.; Popper, B.; Goel, A.; Shakibaei, M. Resveratrol induces chemosensitization to 5-fluorouracil through up-regulation of intercellular junctions, Epithelial-to-mesenchymal transition and apoptosis in colorectal cancer. Biochem. Pharmacol. 2015, 98, 51–68. [Google Scholar] [CrossRef] [PubMed]
  309. Shi, G.; Zheng, X.; Wu, X.; Wang, S.; Wang, Y.; Xing, F. All-trans retinoic acid reverses epithelial-mesenchymal transition in paclitaxel-resistant cells by inhibiting nuclear factor kappa B and upregulating gap junctions. Cancer Sci. 2019, 110, 379–388. [Google Scholar] [CrossRef] [PubMed]
  310. Murphy, S.F.; Varghese, R.T.; Lamouille, S.; Guo, S.; Pridham, K.J.; Kanabur, P.; Osimani, A.M.; Sharma, S.; Jourdan, J.; Rodgers, C.M.; et al. Connexin 43 Inhibition Sensitizes Chemoresistant Glioblastoma Cells to Temozolomide. Cancer Res. 2016, 76, 139–149. [Google Scholar] [CrossRef] [PubMed]
  311. Munoz, J.L.; Rodriguez-Cruz, V.; Greco, S.J.; Ramkissoon, S.H.; Ligon, K.L.; Rameshwar, P. Temozolomide resistance in glioblastoma cells occurs partly through epidermal growth factor receptor-mediated induction of connexin 43. Cell Death Dis. 2014, 5, e1145. [Google Scholar] [CrossRef] [PubMed]
  312. Chen, W.; Wang, D.; Du, X.; He, Y.; Chen, S.; Shao, Q.; Ma, C.; Huang, B.; Chen, A.; Zhao, P.; et al. Glioma cells escaped from cytotoxicity of temozolomide and vincristine by communicating with human astrocytes. Med. Oncol. 2015, 32, 43. [Google Scholar] [CrossRef] [PubMed]
  313. Grek, C.L.; Sheng, Z.; Naus, C.C.; Sin, W.C.; Gourdie, R.G.; Ghatnekar, G.G. Novel approach to temozolomide resistance in malignant glioma: Connexin43-directed therapeutics. Curr. Opin. Pharmacol. 2018, 41, 79–88. [Google Scholar] [CrossRef] [PubMed]
  314. Iyyathurai, J.; Decuypere, J.P.; Leybaert, L.; D’Hondt, C.; Bultynck, G. Connexins: Substrates and regulators of autophagy. BMC Cell Biol. 2016, 17 (Suppl. S1), 20. [Google Scholar] [CrossRef] [PubMed]
  315. Fan, L.X.; Tao, L.; Lai, Y.C.; Cai, S.Y.; Zhao, Z.Y.; Yang, F.; Su, R.Y.; Wang, Q. Cx32 promotes autophagy and produces resistance to SN-induced apoptosis via activation of AMPK signalling in cervical cancer. Int. J. Oncol. 2022, 60, 1–11. [Google Scholar] [CrossRef] [PubMed]
  316. Zhao, Y.; Lai, Y.; Ge, H.; Guo, Y.; Feng, X.; Song, J.; Wang, Q.; Fan, L.; Peng, Y.; Cao, M.; et al. Non-junctional Cx32 mediates anti-apoptotic and pro-tumor effects via epidermal growth factor receptor in human cervical cancer cells. Cell Death Dis. 2017, 8, e2773. [Google Scholar] [CrossRef] [PubMed]
  317. Lampe, P.D.; Laird, D.W. Recent advances in connexin gap junction biology. Fac. Rev. 2022, 11, 14. [Google Scholar] [CrossRef] [PubMed]
  318. Loewenstein, W.R. Junctional intercellular communication: The cell-to-cell membrane channel. Physiol. Rev. 1981, 61, 829–913. [Google Scholar] [CrossRef]
  319. Goldberg, G.S.; Valiunas, V.; Brink, P.R. Selective permeability of gap junction channels. Biochim. Biophys. Acta 2004, 1662, 96–101. [Google Scholar] [CrossRef]
  320. Wei, C.J.; Xu, X.; Lo, C.W. Connexins and cell signaling in development and disease. Annu. Rev. Cell Dev. Biol. 2004, 20, 811–838. [Google Scholar] [CrossRef]
  321. Evans, W.H.; Martin, P.E. Gap junctions: Structure and function (Review). Mol. Membr. Biol. 2002, 19, 121–136. [Google Scholar] [CrossRef]
  322. Spray, D.C.; White, R.L.; Mazet, F.; Bennett, M.V. Regulation of gap junctional conductance. Am. J. Physiol. 1985, 248 Pt 2, H753–H764. [Google Scholar] [CrossRef]
  323. Luo, C.; Yuan, D.; Yao, W.; Cai, J.; Zhou, S.; Zhang, Y.; Hei, Z. Dexmedetomidine protects against apoptosis induced by hypoxia/reoxygenation through the inhibition of gap junctions in NRK-52E cells. Life Sci. 2015, 122, 72–77. [Google Scholar] [CrossRef] [PubMed]
  324. de Pina-Benabou, M.H.; Szostak, V.; Kyrozis, A.; Rempe, D.; Uziel, D.; Urban-Maldonado, M.; Benabou, S.; Spray, D.C.; Federoff, H.J.; Stanton, P.K.; et al. Blockade of gap junctions in vivo provides neuroprotection after perinatal global ischemia. Stroke 2005, 36, 2232–2237. [Google Scholar] [CrossRef] [PubMed]
  325. de Rivero Vaccari, J.C.; Corriveau, R.A.; Belousov, A.B. Gap junctions are required for NMDA receptor dependent cell death in developing neurons. J. Neurophysiol. 2007, 98, 2878–2886. [Google Scholar] [CrossRef] [PubMed]
  326. Peixoto, P.M.; Ryu, S.Y.; Pruzansky, D.P.; Kuriakose, M.; Gilmore, A.; Kinnally, K.W. Mitochondrial apoptosis is amplified through gap junctions. Biochem. Biophys. Res. Commun. 2009, 390, 38–43. [Google Scholar] [CrossRef] [PubMed]
  327. Saito, C.; Shinzawa, K.; Tsujimoto, Y. Synchronized necrotic death of attached hepatocytes mediated via gap junctions. Sci. Rep. 2014, 4, 5169. [Google Scholar] [CrossRef] [PubMed]
  328. Satomi, Y.; Nishino, H.; Shibata, S. Glycyrrhetinic acid and related compounds induce G1 arrest and apoptosis in human hepatocellular carcinoma HepG2. Anticancer Res. 2005, 25, 4043–4047. [Google Scholar] [PubMed]
  329. Yu, J.; Berga, S.L.; Zou, W.; Sun, H.Y.; Johnston-MacAnanny, E.; Yalcinkaya, T.; Sidell, N.; Bagchi, I.C.; Bagchi, M.K.; Taylor, R.N. Gap junction blockade induces apoptosis in human endometrial stromal cells. Mol. Reprod. Dev. 2014, 81, 666–675. [Google Scholar] [CrossRef] [PubMed]
  330. Du, Z.J.; Cui, G.Q.; Zhang, J.; Liu, X.M.; Zhang, Z.H.; Jia, Q.; Ng, J.C.; Peng, C.; Bo, C.X.; Shao, H. Inhibition of gap junction intercellular communication is involved in silica nanoparticles-induced H9c2 cardiomyocytes apoptosis via the mitochondrial pathway. Int. J. Nanomed. 2017, 12, 2179–2188. [Google Scholar] [CrossRef]
  331. Li, C.; Shi, L.; Peng, C.; Yu, G.; Zhang, Y.; Du, Z. Lead-induced cardiomyocytes apoptosis by inhibiting gap junction intercellular communication via autophagy activation. Chem. Biol. Interact. 2021, 337, 109331. [Google Scholar] [CrossRef]
  332. Wang, Q.; Ma, Y.; Li, Y.; He, Z.; Feng, B. Lead-induced cardiomyocytes apoptosis by inhibiting gap junction intercellular communication via modulating the PKCalpha/Cx43 signaling pathway. Chem. Biol. Interact. 2023, 376, 110451. [Google Scholar] [CrossRef] [PubMed]
  333. Kandouz, M.; Batist, G. Gap junctions and connexins as therapeutic targets in cancer. Expert Opin. Ther. Targets 2010, 14, 681–692. [Google Scholar] [CrossRef] [PubMed]
  334. Pitts, J.D. Cancer gene therapy: A bystander effect using the gap junctional pathway. Mol. Carcinog. 1994, 11, 127–130. [Google Scholar] [CrossRef] [PubMed]
  335. Culver, K.W.; Ram, Z.; Wallbridge, S.; Ishii, H.; Oldfield, E.H.; Blaese, R.M. In vivo gene transfer with retroviral vector-producer cells for treatment of experimental brain tumors. Science 1992, 256, 1550–1552. [Google Scholar] [CrossRef] [PubMed]
  336. Takamiya, Y.; Short, M.P.; Ezzeddine, Z.D.; Moolten, F.L.; Breakefield, X.O.; Martuza, R.L. Gene therapy of malignant brain tumors: A rat glioma line bearing the herpes simplex virus type 1-thymidine kinase gene and wild type retrovirus kills other tumor cells. J. Neurosci. Res. 1992, 33, 493–503. [Google Scholar] [CrossRef] [PubMed]
  337. Yahyapour, R.; Motevaseli, E.; Rezaeyan, A.; Abdollahi, H.; Farhood, B.; Cheki, M.; Najafi, M.; Villa, V. Mechanisms of Radiation Bystander and Non-Targeted Effects: Implications to Radiation Carcinogenesis and Radiotherapy. Curr. Radiopharm. 2018, 11, 34–45. [Google Scholar] [CrossRef]
  338. Mukherjee, S.; Chakraborty, A. Radiation-induced bystander phenomenon: Insight and implications in radiotherapy. Int. J. Radiat. Biol. 2019, 95, 243–263. [Google Scholar] [CrossRef]
  339. Prise, K.M.; O’Sullivan, J.M. Radiation-induced bystander signalling in cancer therapy. Nat. Rev. Cancer 2009, 9, 351–360. [Google Scholar] [CrossRef]
  340. Azzam, E.I.; de Toledo, S.M.; Little, J.B. Direct evidence for the participation of gap junction-mediated intercellular communication in the transmission of damage signals from alpha -particle irradiated to nonirradiated cells. Proc. Natl. Acad. Sci. USA 2001, 98, 473–478. [Google Scholar] [CrossRef]
  341. Reid, R.; Mar, E.C.; Huang, E.S.; Topal, M.D. Insertion and extension of acyclic, dideoxy, and ara nucleotides by herpesviridae, human alpha and human beta polymerases. A unique inhibition mechanism for 9-(1,3-dihydroxy-2-propoxymethyl)guanine triphosphate. J. Biol. Chem. 1988, 263, 3898–3904. [Google Scholar] [CrossRef]
  342. Beltinger, C.; Fulda, S.; Kammertoens, T.; Meyer, E.; Uckert, W.; Debatin, K.M. Herpes simplex virus thymidine kinase/ganciclovir-induced apoptosis involves ligand-independent death receptor aggregation and activation of caspases. Proc. Natl. Acad. Sci. USA 1999, 96, 8699–8704. [Google Scholar] [CrossRef] [PubMed]
  343. Morgan, W.F.; Hartmann, A.; Limoli, C.L.; Nagar, S.; Ponnaiya, B. Bystander effects in radiation-induced genomic instability. Mutat. Res. 2002, 504, 91–100. [Google Scholar] [CrossRef] [PubMed]
  344. Klammer, H.; Mladenov, E.; Li, F.; Iliakis, G. Bystander effects as manifestation of intercellular communication of DNA damage and of the cellular oxidative status. Cancer Lett. 2015, 356, 58–71. [Google Scholar] [CrossRef] [PubMed]
  345. Jalal, N.; Haq, S.; Anwar, N.; Nazeer, S.; Saeed, U. Radiation induced bystander effect and DNA damage. J. Cancer Res. Ther. 2014, 10, 819–833. [Google Scholar] [CrossRef] [PubMed]
  346. Tang, H.; Cai, L.; He, X.; Niu, Z.; Huang, H.; Hu, W.; Bian, H.; Huang, H. Radiation-induced bystander effect and its clinical implications. Front. Oncol. 2023, 13, 1124412. [Google Scholar] [CrossRef] [PubMed]
  347. Tartier, L.; Gilchrist, S.; Burdak-Rothkamm, S.; Folkard, M.; Prise, K.M. Cytoplasmic irradiation induces mitochondrial-dependent 53BP1 protein relocalization in irradiated and bystander cells. Cancer Res. 2007, 67, 5872–5879. [Google Scholar] [CrossRef] [PubMed]
  348. Ciccia, A.; Elledge, S.J. The DNA damage response: Making it safe to play with knives. Mol. Cell 2010, 40, 179–204. [Google Scholar] [CrossRef] [PubMed]
  349. Zhou, B.B.; Elledge, S.J. The DNA damage response: Putting checkpoints in perspective. Nature 2000, 408, 433–439. [Google Scholar] [CrossRef]
  350. Batchelor, E.; Mock, C.S.; Bhan, I.; Loewer, A.; Lahav, G. Recurrent initiation: A mechanism for triggering p53 pulses in response to DNA damage. Mol. Cell 2008, 30, 277–289. [Google Scholar] [CrossRef]
  351. Allocati, N.; Di Ilio, C.; De Laurenzi, V. p63/p73 in the control of cell cycle and cell death. Exp. Cell Res. 2012, 318, 1285–1290. [Google Scholar] [CrossRef]
  352. Matt, S.; Hofmann, T.G. The DNA damage-induced cell death response: A roadmap to kill cancer cells. Cell Mol. Life Sci. 2016, 73, 2829–2850. [Google Scholar] [CrossRef]
  353. Shao, C.; Folkard, M.; Michael, B.D.; Prise, K.M. Targeted cytoplasmic irradiation induces bystander responses. Proc. Natl. Acad. Sci. USA 2004, 101, 13495–13500. [Google Scholar] [CrossRef] [PubMed]
  354. Wu, L.J.; Randers-Pehrson, G.; Xu, A.; Waldren, C.A.; Geard, C.R.; Yu, Z.; Hei, T.K. Targeted cytoplasmic irradiation with alpha particles induces mutations in mammalian cells. Proc. Natl. Acad. Sci. USA 1999, 96, 4959–4964. [Google Scholar] [CrossRef] [PubMed]
  355. Narayanan, P.K.; Goodwin, E.H.; Lehnert, B.E. Alpha particles initiate biological production of superoxide anions and hydrogen peroxide in human cells. Cancer Res. 1997, 57, 3963–3971. [Google Scholar] [PubMed]
  356. Nakagami, Y.; Ito, M.; Hara, T.; Inoue, T.; Matsubara, S. Nuclear translocation of DNase II and acid phosphatase during radiation-induced apoptosis in HL60 cells. Acta Oncol. 2003, 42, 227–236. [Google Scholar] [CrossRef] [PubMed]
  357. Vit, J.P.; Rosselli, F. Role of the ceramide-signaling pathways in ionizing radiation-induced apoptosis. Oncogene 2003, 22, 8645–8652. [Google Scholar] [CrossRef] [PubMed]
  358. Santana, P.; Pena, L.A.; Haimovitz-Friedman, A.; Martin, S.; Green, D.; McLoughlin, M.; Cordon-Cardo, C.; Schuchman, E.H.; Fuks, Z.; Kolesnick, R. Acid sphingomyelinase-deficient human lymphoblasts and mice are defective in radiation-induced apoptosis. Cell 1996, 86, 189–199. [Google Scholar] [CrossRef]
  359. Rebillard, A.; Rioux-Leclercq, N.; Muller, C.; Bellaud, P.; Jouan, F.; Meurette, O.; Jouan, E.; Vernhet, L.; Le Quement, C.; Carpinteiro, A.; et al. Acid sphingomyelinase deficiency protects from cisplatin-induced gastrointestinal damage. Oncogene 2008, 27, 6590–6595. [Google Scholar] [CrossRef]
  360. Smith, E.L.; Schuchman, E.H. The unexpected role of acid sphingomyelinase in cell death and the pathophysiology of common diseases. FASEB J. 2008, 22, 3419–3431. [Google Scholar] [CrossRef]
  361. Prokhorova, E.A.; Egorshina, A.Y.; Zhivotovsky, B.; Kopeina, G.S. The DNA-damage response and nuclear events as regulators of nonapoptotic forms of cell death. Oncogene 2020, 39, 1–16. [Google Scholar] [CrossRef]
  362. Vitale, I.; Galluzzi, L.; Castedo, M.; Kroemer, G. Mitotic catastrophe: A mechanism for avoiding genomic instability. Nat. Rev. Mol. Cell Biol. 2011, 12, 385–392. [Google Scholar] [CrossRef]
  363. Apel, A.; Herr, I.; Schwarz, H.; Rodemann, H.P.; Mayer, A. Blocked autophagy sensitizes resistant carcinoma cells to radiation therapy. Cancer Res. 2008, 68, 1485–1494. [Google Scholar] [CrossRef] [PubMed]
  364. Gewirtz, D.A.; Hilliker, M.L.; Wilson, E.N. Promotion of autophagy as a mechanism for radiation sensitization of breast tumor cells. Radiother. Oncol. 2009, 92, 323–328. [Google Scholar] [CrossRef]
  365. Makowska, A.; Eble, M.; Prescher, K.; Hoss, M.; Kontny, U. Chloroquine Sensitizes Nasopharyngeal Carcinoma Cells but Not Nasoepithelial Cells to Irradiation by Blocking Autophagy. PLoS ONE 2016, 11, e0166766. [Google Scholar] [CrossRef] [PubMed]
  366. Palumbo, S.; Comincini, S. Autophagy and ionizing radiation in tumors: The “survive or not survive” dilemma. J. Cell Physiol. 2013, 228, 1–8. [Google Scholar] [CrossRef] [PubMed]
  367. Hino, M.; Hamada, N.; Tajika, Y.; Funayama, T.; Morimura, Y.; Sakashita, T.; Yokota, Y.; Fukamoto, K.; Mutou, Y.; Kobayashi, Y.; et al. Heavy ion irradiation induces autophagy in irradiated C2C12 myoblasts and their bystander cells. J. Electron. Microsc. 2010, 59, 495–501. [Google Scholar] [CrossRef] [PubMed]
  368. Wang, X.; Zhang, J.; Fu, J.; Wang, J.; Ye, S.; Liu, W.; Shao, C. Role of ROS-mediated autophagy in radiation-induced bystander effect of hepatoma cells. Int. J. Radiat. Biol. 2015, 91, 452–458. [Google Scholar] [CrossRef] [PubMed]
  369. Mathew, R.; Kongara, S.; Beaudoin, B.; Karp, C.M.; Bray, K.; Degenhardt, K.; Chen, G.; Jin, S.; White, E. Autophagy suppresses tumor progression by limiting chromosomal instability. Genes Dev. 2007, 21, 1367–1381. [Google Scholar] [CrossRef]
  370. Karantza-Wadsworth, V.; Patel, S.; Kravchuk, O.; Chen, G.; Mathew, R.; Jin, S.; White, E. Autophagy mitigates metabolic stress and genome damage in mammary tumorigenesis. Genes Dev. 2007, 21, 1621–1635. [Google Scholar] [CrossRef]
  371. Ren, Y.; Yang, P.; Li, C.; Wang, W.A.; Zhang, T.; Li, J.; Li, H.; Dong, C.; Meng, W.; Zhou, H. Ionizing radiation triggers mitophagy to enhance DNA damage in cancer cells. Cell Death Discov. 2023, 9, 267. [Google Scholar] [CrossRef]
  372. Yang, P.; Li, J.; Zhang, T.; Ren, Y.; Zhang, Q.; Liu, R.; Li, H.; Hua, J.; Wang, W.A.; Wang, J.; et al. Ionizing radiation-induced mitophagy promotes ferroptosis by increasing intracellular free fatty acids. Cell Death Differ. 2023, 30, 2432–2445. [Google Scholar] [CrossRef] [PubMed]
  373. Dixon, S.J.; Lemberg, K.M.; Lamprecht, M.R.; Skouta, R.; Zaitsev, E.M.; Gleason, C.E.; Patel, D.N.; Bauer, A.J.; Cantley, A.M.; Yang, W.S.; et al. Ferroptosis: An iron-dependent form of nonapoptotic cell death. Cell 2012, 149, 1060–1072. [Google Scholar] [CrossRef] [PubMed]
  374. Chen, X.; Li, J.; Kang, R.; Klionsky, D.J.; Tang, D. Ferroptosis: Machinery and regulation. Autophagy 2021, 17, 2054–2081. [Google Scholar] [CrossRef] [PubMed]
  375. 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]
  376. Zheng, J.; Conrad, M. The Metabolic Underpinnings of Ferroptosis. Cell Metab. 2020, 32, 920–937. [Google Scholar] [CrossRef] [PubMed]
  377. Liang, D.; Minikes, A.M.; Jiang, X. Ferroptosis at the intersection of lipid metabolism and cellular signaling. Mol. Cell 2022, 82, 2215–2227. [Google Scholar] [CrossRef] [PubMed]
  378. Pope, L.E.; Dixon, S.J. Regulation of ferroptosis by lipid metabolism. Trends Cell Biol. 2023, 33, 1077–1087. [Google Scholar] [CrossRef]
  379. Yang, W.S.; Stockwell, B.R. Ferroptosis: Death by Lipid Peroxidation. Trends Cell Biol. 2016, 26, 165–176. [Google Scholar] [CrossRef]
  380. Naowarojna, N.; Wu, T.W.; Pan, Z.; Li, M.; Han, J.R.; Zou, Y. Dynamic Regulation of Ferroptosis by Lipid Metabolism. Antioxid. Redox Signal. 2023, 39, 59–78. [Google Scholar] [CrossRef]
  381. Zhang, X.; Li, X.; Zheng, C.; Yang, C.; Zhang, R.; Wang, A.; Feng, J.; Hu, X.; Chang, S.; Zhang, H. Ferroptosis, a new form of cell death defined after radiation exposure. Int. J. Radiat. Biol. 2022, 98, 1201–1209. [Google Scholar] [CrossRef]
  382. Shibata, Y.; Yasui, H.; Higashikawa, K.; Miyamoto, N.; Kuge, Y. Erastin, a ferroptosis-inducing agent, sensitized cancer cells to X-ray irradiation via glutathione starvation in vitro and in vivo. PLoS ONE 2019, 14, e0225931. [Google Scholar] [CrossRef]
  383. Riegman, M.; Sagie, L.; Galed, C.; Levin, T.; Steinberg, N.; Dixon, S.J.; Wiesner, U.; Bradbury, M.S.; Niethammer, P.; Zaritsky, A.; et al. Ferroptosis occurs through an osmotic mechanism and propagates independently of cell rupture. Nat. Cell Biol. 2020, 22, 1042–1048. [Google Scholar] [CrossRef]
  384. Jelinek, A.; Heyder, L.; Daude, M.; Plessner, M.; Krippner, S.; Grosse, R.; Diederich, W.E.; Culmsee, C. Mitochondrial rescue prevents glutathione peroxidase-dependent ferroptosis. Free Radic. Biol. Med. 2018, 117, 45–57. [Google Scholar] [CrossRef] [PubMed]
  385. Gao, M.; Yi, J.; Zhu, J.; Minikes, A.M.; Monian, P.; Thompson, C.B.; Jiang, X. Role of Mitochondria in Ferroptosis. Mol. Cell 2019, 73, 354–363. [Google Scholar] [CrossRef] [PubMed]
  386. Gan, B. Mitochondrial regulation of ferroptosis. J. Cell Biol. 2021, 220, e202105043. [Google Scholar] [CrossRef] [PubMed]
  387. Huang, Q.; Sha, W.; Gu, Q.; Wang, J.; Zhu, Y.; Xu, T.; Xu, Z.; Yan, F.; Lin, X.; Tian, S. Inhibition of Connexin43 Improves the Recovery of Spinal Cord Injury Against Ferroptosis via the SLC7A11/GPX4 Pathway. Neuroscience 2023, 526, 121–134. [Google Scholar] [CrossRef] [PubMed]
  388. Losa, D.; Kohler, T.; Bellec, J.; Dudez, T.; Crespin, S.; Bacchetta, M.; Boulanger, P.; Hong, S.S.; Morel, S.; Nguyen, T.H.; et al. Pseudomonas aeruginosa-induced apoptosis in airway epithelial cells is mediated by gap junctional communication in a JNK-dependent manner. J. Immunol. 2014, 192, 4804–4812. [Google Scholar] [CrossRef] [PubMed]
  389. Su, L.; Jiang, X.; Yang, C.; Zhang, J.; Chen, B.; Li, Y.; Yao, S.; Xie, Q.; Gomez, H.; Murugan, R.; et al. Pannexin 1 mediates ferroptosis that contributes to renal ischemia/reperfusion injury. J. Biol. Chem. 2019, 294, 19395–19404. [Google Scholar] [CrossRef]
  390. Pelegrin, P.; Surprenant, A. Pannexin-1 mediates large pore formation and interleukin-1beta release by the ATP-gated P2X7 receptor. EMBO J. 2006, 25, 5071–5082. [Google Scholar] [CrossRef]
  391. Iglesias, R.; Dahl, G.; Qiu, F.; Spray, D.C.; Scemes, E. Pannexin 1: The molecular substrate of astrocyte “hemichannels”. J. Neurosci. 2009, 29, 7092–7097. [Google Scholar] [CrossRef]
  392. Bruzzone, R.; Hormuzdi, S.G.; Barbe, M.T.; Herb, A.; Monyer, H. Pannexins, a family of gap junction proteins expressed in brain. Proc. Natl. Acad. Sci. USA 2003, 100, 13644–13649. [Google Scholar] [CrossRef]
  393. Linkermann, A.; Skouta, R.; Himmerkus, N.; Mulay, S.R.; Dewitz, C.; De Zen, F.; Prokai, A.; Zuchtriegel, G.; Krombach, F.; Welz, P.S.; et al. Synchronized renal tubular cell death involves ferroptosis. Proc. Natl. Acad. Sci. USA 2014, 111, 16836–16841. [Google Scholar] [CrossRef]
  394. Li, X.; Peng, X.; Zhou, X.; Li, M.; Chen, G.; Shi, W.; Yu, H.; Zhang, C.; Li, Y.; Feng, Z.; et al. Small extracellular vesicles delivering lncRNA WAC-AS1 aggravate renal allograft ischemia-reperfusion injury by inducing ferroptosis propagation. Cell Death Differ. 2023, 30, 2167–2186. [Google Scholar] [CrossRef] [PubMed]
  395. Kawasaki, N.K.; Suhara, T.; Komai, K.; Shimada, B.K.; Yorichika, N.; Kobayashi, M.; Baba, Y.; Higa, J.K.; Matsui, T. The role of ferroptosis in cell-to-cell propagation of cell death initiated from focal injury in cardiomyocytes. Life Sci. 2023, 332, 122113. [Google Scholar] [CrossRef] [PubMed]
  396. Zampieri, L.X.; Silva-Almeida, C.; Rondeau, J.D.; Sonveaux, P. Mitochondrial Transfer in Cancer: A Comprehensive Review. Int. J. Mol. Sci. 2021, 22, 3245. [Google Scholar] [CrossRef] [PubMed]
  397. Torralba, D.; Baixauli, F.; Sanchez-Madrid, F. Mitochondria Know No Boundaries: Mechanisms and Functions of Intercellular Mitochondrial Transfer. Front. Cell Dev. Biol. 2016, 4, 107. [Google Scholar] [CrossRef] [PubMed]
  398. Rustom, A.; Saffrich, R.; Markovic, I.; Walther, P.; Gerdes, H.H. Nanotubular highways for intercellular organelle transport. Science 2004, 303, 1007–1010. [Google Scholar] [CrossRef] [PubMed]
  399. Norris, R.P. Transfer of mitochondria and endosomes between cells by gap junction internalization. Traffic 2021, 22, 174–179. [Google Scholar] [CrossRef] [PubMed]
  400. Rogers, R.S.; Bhattacharya, J. When cells become organelle donors. Physiology 2013, 28, 414–422. [Google Scholar] [CrossRef]
  401. Murray, L.M.A.; Krasnodembskaya, A.D. Concise Review: Intercellular Communication Via Organelle Transfer in the Biology and Therapeutic Applications of Stem Cells. Stem Cells 2019, 37, 14–25. [Google Scholar] [CrossRef]
  402. Hayakawa, K.; Esposito, E.; Wang, X.; Terasaki, Y.; Liu, Y.; Xing, C.; Ji, X.; Lo, E.H. Transfer of mitochondria from astrocytes to neurons after stroke. Nature 2016, 535, 551–555. [Google Scholar] [CrossRef] [PubMed]
  403. Joshi, A.U.; Minhas, P.S.; Liddelow, S.A.; Haileselassie, B.; Andreasson, K.I.; Dorn, G.W., 2nd; Mochly-Rosen, D. Fragmented mitochondria released from microglia trigger A1 astrocytic response and propagate inflammatory neurodegeneration. Nat. Neurosci. 2019, 22, 1635–1648. [Google Scholar] [CrossRef] [PubMed]
  404. Peruzzotti-Jametti, L.; Bernstock, J.D.; Willis, C.M.; Manferrari, G.; Rogall, R.; Fernandez-Vizarra, E.; Williamson, J.C.; Braga, A.; van den Bosch, A.; Leonardi, T.; et al. Neural stem cells traffic functional mitochondria via extracellular vesicles. PLoS Biol. 2021, 19, e3001166. [Google Scholar] [CrossRef]
  405. Jin, S.; Cordes, N. ATM controls DNA repair and mitochondria transfer between neighboring cells. Cell Commun. Signal. 2019, 17, 144. [Google Scholar] [CrossRef] [PubMed]
  406. Watson, D.C.; Bayik, D.; Storevik, S.; Moreino, S.S.; Sprowls, S.A.; Han, J.; Augustsson, M.T.; Lauko, A.; Sravya, P.; Rosland, G.V.; et al. GAP43-dependent mitochondria transfer from astrocytes enhances glioblastoma tumorigenicity. Nat. Cancer 2023, 4, 648–664. [Google Scholar] [CrossRef]
  407. Walsh, D.W.M.; Siebenwirth, C.; Greubel, C.; Ilicic, K.; Reindl, J.; Girst, S.; Muggiolu, G.; Simon, M.; Barberet, P.; Seznec, H.; et al. Live cell imaging of mitochondria following targeted irradiation in situ reveals rapid and highly localized loss of membrane potential. Sci. Rep. 2017, 7, 46684. [Google Scholar] [CrossRef] [PubMed]
  408. Yoon, J.; Ryu, S.W.; Lee, S.; Choi, C. Cytosolic irradiation of femtosecond laser induces mitochondria-dependent apoptosis-like cell death via intrinsic reactive oxygen cascades. Sci. Rep. 2015, 5, 8231. [Google Scholar] [CrossRef]
  409. Santavanond, J.P.; Rutter, S.F.; Atkin-Smith, G.K.; Poon, I.K.H. Apoptotic Bodies: Mechanism of Formation, Isolation and Functional Relevance. Subcell. Biochem. 2021, 97, 61–88. [Google Scholar] [CrossRef]
  410. Thery, C.; Witwer, K.W.; Aikawa, E.; Alcaraz, M.J.; Anderson, J.D.; Andriantsitohaina, R.; Antoniou, A.; Arab, T.; Archer, F.; Atkin-Smith, G.K.; et al. Minimal information for studies of extracellular vesicles 2018 (MISEV2018): A position statement of the International Society for Extracellular Vesicles and update of the MISEV2014 guidelines. J. Extracell. Vesicles 2018, 7, 1535750. [Google Scholar] [CrossRef]
  411. Gyorgy, B.; Szabo, T.G.; Pasztoi, M.; Pal, Z.; Misjak, P.; Aradi, B.; Laszlo, V.; Pallinger, E.; Pap, E.; Kittel, A.; et al. Membrane vesicles, current state-of-the-art: Emerging role of extracellular vesicles. Cell Mol. Life Sci. 2011, 68, 2667–2688. [Google Scholar] [CrossRef]
  412. Tixeira, R.; Caruso, S.; Paone, S.; Baxter, A.A.; Atkin-Smith, G.K.; Hulett, M.D.; Poon, I.K. Defining the morphologic features and products of cell disassembly during apoptosis. Apoptosis 2017, 22, 475–477. [Google Scholar] [CrossRef]
  413. Zhang, Y.; Chen, X.; Gueydan, C.; Han, J. Plasma membrane changes during programmed cell deaths. Cell Res. 2018, 28, 9–21. [Google Scholar] [CrossRef] [PubMed]
  414. Schiller, M.; Bekeredjian-Ding, I.; Heyder, P.; Blank, N.; Ho, A.D.; Lorenz, H.M. Autoantigens are translocated into small apoptotic bodies during early stages of apoptosis. Cell Death Differ. 2008, 15, 183–191. [Google Scholar] [CrossRef] [PubMed]
  415. Maas, S.L.N.; Breakefield, X.O.; Weaver, A.M. Extracellular Vesicles: Unique Intercellular Delivery Vehicles. Trends Cell Biol. 2017, 27, 172–188. [Google Scholar] [CrossRef]
  416. Thery, C.; Zitvogel, L.; Amigorena, S. Exosomes: Composition, biogenesis and function. Nat. Rev. Immunol. 2002, 2, 569–579. [Google Scholar] [CrossRef] [PubMed]
  417. Nieuwland, R.; Falcon-Perez, J.M.; Soekmadji, C.; Boilard, E.; Carter, D.; Buzas, E.I. Essentials of extracellular vesicles: Posters on basic and clinical aspects of extracellular vesicles. J. Extracell. Vesicles 2018, 7, 1548234. [Google Scholar] [CrossRef] [PubMed]
  418. Tricarico, C.; Clancy, J.; D’Souza-Schorey, C. Biology and biogenesis of shed microvesicles. Small GTPases 2017, 8, 220–232. [Google Scholar] [CrossRef]
  419. Colombo, M.; Moita, C.; van Niel, G.; Kowal, J.; Vigneron, J.; Benaroch, P.; Manel, N.; Moita, L.F.; Thery, C.; Raposo, G. Analysis of ESCRT functions in exosome biogenesis, composition and secretion highlights the heterogeneity of extracellular vesicles. J. Cell Sci. 2013, 126 Pt 24, 5553–5565. [Google Scholar] [CrossRef]
  420. Guescini, M.; Guidolin, D.; Vallorani, L.; Casadei, L.; Gioacchini, A.M.; Tibollo, P.; Battistelli, M.; Falcieri, E.; Battistin, L.; Agnati, L.F.; et al. C2C12 myoblasts release micro-vesicles containing mtDNA and proteins involved in signal transduction. Exp. Cell Res. 2010, 316, 1977–1984. [Google Scholar] [CrossRef]
  421. Loyer, X.; Vion, A.C.; Tedgui, A.; Boulanger, C.M. Microvesicles as cell-cell messengers in cardiovascular diseases. Circ. Res. 2014, 114, 345–353. [Google Scholar] [CrossRef]
  422. Simons, M.; Raposo, G. Exosomes—Vesicular carriers for intercellular communication. Curr. Opin. Cell Biol. 2009, 21, 575–581. [Google Scholar] [CrossRef]
  423. Ludwig, A.K.; Giebel, B. Exosomes: Small vesicles participating in intercellular communication. Int. J. Biochem. Cell Biol. 2012, 44, 11–15. [Google Scholar] [CrossRef] [PubMed]
  424. Lopatina, T.; Gai, C.; Deregibus, M.C.; Kholia, S.; Camussi, G. Cross Talk between Cancer and Mesenchymal Stem Cells through Extracellular Vesicles Carrying Nucleic Acids. Front. Oncol. 2016, 6, 125. [Google Scholar] [CrossRef] [PubMed]
  425. Fruhbeis, C.; Frohlich, D.; Kuo, W.P.; Kramer-Albers, E.M. Extracellular vesicles as mediators of neuron-glia communication. Front. Cell Neurosci. 2013, 7, 182. [Google Scholar] [CrossRef] [PubMed]
  426. Wendler, F.; Bota-Rabassedas, N.; Franch-Marro, X. Cancer becomes wasteful: Emerging roles of exosomes(dagger) in cell-fate determination. J. Extracell. Vesicles 2013, 2, 22390. [Google Scholar] [CrossRef] [PubMed]
  427. Wolfers, J.; Lozier, A.; Raposo, G.; Regnault, A.; Thery, C.; Masurier, C.; Flament, C.; Pouzieux, S.; Faure, F.; Tursz, T.; et al. Tumor-derived exosomes are a source of shared tumor rejection antigens for CTL cross-priming. Nat. Med. 2001, 7, 297–303. [Google Scholar] [CrossRef]
  428. Thery, C.; Ostrowski, M.; Segura, E. Membrane vesicles as conveyors of immune responses. Nat. Rev. Immunol. 2009, 9, 581–593. [Google Scholar] [CrossRef] [PubMed]
  429. Raposo, G.; Stoorvogel, W. Extracellular vesicles: Exosomes, microvesicles, and friends. J. Cell Biol. 2013, 200, 373–383. [Google Scholar] [CrossRef] [PubMed]
  430. Taylor, D.D.; Gercel-Taylor, C. Tumour-derived exosomes and their role in cancer-associated T-cell signalling defects. Br. J. Cancer 2005, 92, 305–311. [Google Scholar] [CrossRef]
  431. Thery, C.; Regnault, A.; Garin, J.; Wolfers, J.; Zitvogel, L.; Ricciardi-Castagnoli, P.; Raposo, G.; Amigorena, S. Molecular characterization of dendritic cell-derived exosomes. Selective accumulation of the heat shock protein hsc73. J. Cell Biol. 1999, 147, 599–610. [Google Scholar] [CrossRef]
  432. Wilson, C.M.; Naves, T.; Vincent, F.; Melloni, B.; Bonnaud, F.; Lalloue, F.; Jauberteau, M.O. Sortilin mediates the release and transfer of exosomes in concert with two tyrosine kinase receptors. J. Cell Sci. 2014, 127, 3983–3997. [Google Scholar] [CrossRef] [PubMed]
  433. Abels, E.R.; Breakefield, X.O. Introduction to Extracellular Vesicles: Biogenesis, RNA Cargo Selection, Content, Release, and Uptake. Cell Mol. Neurobiol. 2016, 36, 301–312. [Google Scholar] [CrossRef] [PubMed]
  434. Beer, L.; Zimmermann, M.; Mitterbauer, A.; Ellinger, A.; Gruber, F.; Narzt, M.S.; Zellner, M.; Gyongyosi, M.; Madlener, S.; Simader, E.; et al. Analysis of the Secretome of Apoptotic Peripheral Blood Mononuclear Cells: Impact of Released Proteins and Exosomes for Tissue Regeneration. Sci. Rep. 2015, 5, 16662. [Google Scholar] [CrossRef]
  435. Du, Y.; Du, S.; Liu, L.; Gan, F.; Jiang, X.; Wangrao, K.; Lyu, P.; Gong, P.; Yao, Y. Radiation-Induced Bystander Effect can be Transmitted Through Exosomes Using miRNAs as Effector Molecules. Radiat. Res. 2020, 194, 89–100. [Google Scholar] [CrossRef] [PubMed]
  436. Smolarz, M.; Skoczylas, L.; Gawin, M.; Krzyzowska, M.; Pietrowska, M.; Widlak, P. Radiation-Induced Bystander Effect Mediated by Exosomes Involves the Replication Stress in Recipient Cells. Int. J. Mol. Sci. 2022, 23, 4169. [Google Scholar] [CrossRef] [PubMed]
  437. Al-Mayah, A.; Bright, S.; Chapman, K.; Irons, S.; Luo, P.; Carter, D.; Goodwin, E.; Kadhim, M. The non-targeted effects of radiation are perpetuated by exosomes. Mutat. Res. 2015, 772, 38–45. [Google Scholar] [CrossRef]
  438. Jella, K.K.; Rani, S.; O’Driscoll, L.; McClean, B.; Byrne, H.J.; Lyng, F.M. Exosomes are involved in mediating radiation induced bystander signaling in human keratinocyte cells. Radiat. Res. 2014, 181, 138–145. [Google Scholar] [CrossRef]
  439. Al-Mayah, A.H.; Irons, S.L.; Pink, R.C.; Carter, D.R.; Kadhim, M.A. Possible role of exosomes containing RNA in mediating nontargeted effect of ionizing radiation. Radiat. Res. 2012, 177, 539–545. [Google Scholar] [CrossRef]
  440. Wang, J.; Ma, W.; Si, C.; Zhang, M.; Qian, W.; Park, G.; Zhou, B.; Luo, D. Exosome-mediated miR-4655-3p contributes to UV radiation-induced bystander effects. Exp. Cell Res. 2022, 418, 113247. [Google Scholar] [CrossRef]
  441. Song, M.; Wang, Y.; Shang, Z.F.; Liu, X.D.; Xie, D.F.; Wang, Q.; Guan, H.; Zhou, P.K. Bystander autophagy mediated by radiation-induced exosomal miR-7-5p in non-targeted human bronchial epithelial cells. Sci. Rep. 2016, 6, 30165. [Google Scholar] [CrossRef]
  442. Ni, N.; Ma, W.; Tao, Y.; Liu, J.; Hua, H.; Cheng, J.; Wang, J.; Zhou, B.; Luo, D. Exosomal MiR-769-5p Exacerbates Ultraviolet-Induced Bystander Effect by Targeting TGFBR1. Front. Physiol. 2020, 11, 603081. [Google Scholar] [CrossRef]
  443. Hu, W.; Xu, S.; Yao, B.; Hong, M.; Wu, X.; Pei, H.; Chang, L.; Ding, N.; Gao, X.; Ye, C.; et al. MiR-663 inhibits radiation-induced bystander effects by targeting TGFB1 in a feedback mode. RNA Biol. 2014, 11, 1189–1198. [Google Scholar] [CrossRef] [PubMed]
  444. Shao, C.; Prise, K.M.; Folkard, M. Signaling factors for irradiated glioma cells induced bystander responses in fibroblasts. Mutat. Res. 2008, 638, 139–145. [Google Scholar] [CrossRef] [PubMed]
  445. Cocca, B.A.; Cline, A.M.; Radic, M.Z. Blebs and apoptotic bodies are B cell autoantigens. J. Immunol. 2002, 169, 159–166. [Google Scholar] [CrossRef] [PubMed]
  446. Segura, E.; Guerin, C.; Hogg, N.; Amigorena, S.; Thery, C. CD8+ dendritic cells use LFA-1 to capture MHC-peptide complexes from exosomes in vivo. J. Immunol. 2007, 179, 1489–1496. [Google Scholar] [CrossRef] [PubMed]
  447. Thery, C.; Duban, L.; Segura, E.; Veron, P.; Lantz, O.; Amigorena, S. Indirect activation of naive CD4+ T cells by dendritic cell-derived exosomes. Nat. Immunol. 2002, 3, 1156–1162. [Google Scholar] [CrossRef] [PubMed]
  448. Martin, S.; Tesse, A.; Hugel, B.; Martinez, M.C.; Morel, O.; Freyssinet, J.M.; Andriantsitohaina, R. Shed membrane particles from T lymphocytes impair endothelial function and regulate endothelial protein expression. Circulation 2004, 109, 1653–1659. [Google Scholar] [CrossRef] [PubMed]
  449. Murray, P.J. Macrophage Polarization. Annu. Rev. Physiol. 2017, 79, 541–566. [Google Scholar] [CrossRef]
  450. Ma, Q.; Liang, M.; Wu, Y.; Luo, F.; Ma, Z.; Dong, S.; Xu, J.; Dou, C. Osteoclast-derived apoptotic bodies couple bone resorption and formation in bone remodeling. Bone Res. 2021, 9, 5. [Google Scholar] [CrossRef]
  451. Brock, C.K.; Wallin, S.T.; Ruiz, O.E.; Samms, K.M.; Mandal, A.; Sumner, E.A.; Eisenhoffer, G.T. Stem cell proliferation is induced by apoptotic bodies from dying cells during epithelial tissue maintenance. Nat. Commun. 2019, 10, 1044. [Google Scholar] [CrossRef]
  452. Zhou, M.; Li, Y.J.; Tang, Y.C.; Hao, X.Y.; Xu, W.J.; Xiang, D.X.; Wu, J.Y. Apoptotic bodies for advanced drug delivery and therapy. J. Control. Release 2022, 351, 394–406. [Google Scholar] [CrossRef]
  453. Kalra, H.; Drummen, G.P.; Mathivanan, S. Focus on Extracellular Vesicles: Introducing the Next Small Big Thing. Int. J. Mol. Sci. 2016, 17, 170. [Google Scholar] [CrossRef] [PubMed]
  454. Freeman, S.M.; Abboud, C.N.; Whartenby, K.A.; Packman, C.H.; Koeplin, D.S.; Moolten, F.L.; Abraham, G.N. The “bystander effect”: Tumor regression when a fraction of the tumor mass is genetically modified. Cancer Res. 1993, 53, 5274–5283. [Google Scholar] [PubMed]
  455. Matejka, N.; Reindl, J. Influence of alpha-Particle Radiation on Intercellular Communication Networks of Tunneling Nanotubes in U87 Glioblastoma Cells. Front. Oncol. 2020, 10, 1691. [Google Scholar] [CrossRef] [PubMed]
  456. Heeran, A.B.; Berrigan, H.P.; O’Sullivan, J. The Radiation-Induced Bystander Effect (RIBE) and its Connections with the Hallmarks of Cancer. Radiat. Res. 2019, 192, 668–679. [Google Scholar] [CrossRef]
  457. Mancuso, M.; Pasquali, E.; Giardullo, P.; Leonardi, S.; Tanori, M.; Di Majo, V.; Pazzaglia, S.; Saran, A. The radiation bystander effect and its potential implications for human health. Curr. Mol. Med. 2012, 12, 613–624. [Google Scholar] [CrossRef] [PubMed]
  458. Kandouz, M. (Ed.) Intercellular Communication in Cancer; Springer: Dordrecht, The Netherlands, 2015. [Google Scholar]
  459. Nagata, S.; Hanayama, R.; Kawane, K. Autoimmunity and the clearance of dead cells. Cell 2010, 140, 619–630. [Google Scholar] [CrossRef]
  460. Kawamoto, Y.; Nakajima, Y.I.; Kuranaga, E. Apoptosis in Cellular Society: Communication between Apoptotic Cells and Their Neighbors. Int. J. Mol. Sci. 2016, 17, 2144. [Google Scholar] [CrossRef]
  461. Reddien, P.W.; Cameron, S.; Horvitz, H.R. Phagocytosis promotes programmed cell death in C. elegans. Nature 2001, 412, 198–202. [Google Scholar] [CrossRef]
  462. Hoeppner, D.J.; Hengartner, M.O.; Schnabel, R. Engulfment genes cooperate with ced-3 to promote cell death in Caenorhabditis elegans. Nature 2001, 412, 202–206. [Google Scholar] [CrossRef]
  463. Chakraborty, S.; Lambie, E.J.; Bindu, S.; Mikeladze-Dvali, T.; Conradt, B. Engulfment pathways promote programmed cell death by enhancing the unequal segregation of apoptotic potential. Nat. Commun. 2015, 6, 10126. [Google Scholar] [CrossRef]
Figure 1. The boundaries between different modes of cell death are not clear-cut, as one often leads to the other, depends on it, or at least shares common features with it.
Figure 1. The boundaries between different modes of cell death are not clear-cut, as one often leads to the other, depends on it, or at least shares common features with it.
Cells 13 00325 g001
Figure 2. Yearly evolution of publication numbers using four research queries related to cell death, as collated from the NCBI PubMed database.
Figure 2. Yearly evolution of publication numbers using four research queries related to cell death, as collated from the NCBI PubMed database.
Cells 13 00325 g002
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

Kandouz, M. Cell Death, by Any Other Name…. Cells 2024, 13, 325. https://doi.org/10.3390/cells13040325

AMA Style

Kandouz M. Cell Death, by Any Other Name…. Cells. 2024; 13(4):325. https://doi.org/10.3390/cells13040325

Chicago/Turabian Style

Kandouz, Mustapha. 2024. "Cell Death, by Any Other Name…" Cells 13, no. 4: 325. https://doi.org/10.3390/cells13040325

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop