IMR Press / JIN / Volume 21 / Issue 1 / DOI: 10.31083/j.jin2101030
Open Access Review
Programmed cell death in cerebellar Purkinje neurons
Show Less
1 Department of Anatomy, Faculty of Medicine, Jordan University of Science and Technology, 22110 Irbid, Jordan
*Correspondence: nserekat@just.edu.jo (Nour S. Erekat)
J. Integr. Neurosci. 2022, 21(1), 30; https://doi.org/10.31083/j.jin2101030
Submitted: 20 April 2021 | Revised: 31 May 2021 | Accepted: 10 August 2021 | Published: 28 January 2022
Copyright: © 2022 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

Apoptosis, autophagy and necrosis are the three main types of programmed cell death. One or more of these types of programmed cell death may take place in neurons leading to their death in various neurodegenerative disorders in humans. Purkinje neurons (PNs) are among the most highly vulnerable population of neurons to cell death in response to intrinsic hereditary diseases or extrinsic toxic, hypoxic, ischemic, and traumatic injury. In this review, we will describe the three main types of programmed cell death, including the molecular mechanisms and the sequence of events in each of them, and thus illustrating the intracellular proteins that mediate and regulate each of these types. Then, we will discuss the role of Ca2+ in PN function and increased vulnerability to cell death. Additionally, PN death will be described in animal models, namely lurcher mutant mouse and shaker mutant rat, in order to illustrate the potential therapeutic implications of programmed cell death in PNs by reviewing the previous studies that were carried out to interfere with the programmed cell death in an attempt to rescue PNs from death.

Keywords
Neurodegeneration
Purkinje neurons
Programmed cell death
Apoptosis
Autophagy
Necrosis
Mutations
Necroptosis
1. Introduction

Neurodegenerative disorders result from the progressive loss of particularly susceptible populations of neurons [1]. Neurodegenerative disorders include Parkinson disease (PD), Alzheimer disease (AD), Huntington disease (HD), and ataxias [2, 3]. Neurons die during development, aging and consequent to a number of pathological factors, such as genetic mutations and traumatic injury.

Neuronal death can occur by one or more of three chief programmed death pathways, which are apoptosis, autophagy, and necrosis [4, 5, 6]. Apoptosis and autophagy are referred to as programmed cell death type I and type II, respectively [7]. Necrosis had been considered as a non-programmed cell death for a long time. However, necrosis has more recently been included as a type of programmed cell death [7, 8]. Thus, this review will provide an overview on the chief types of programmed cell death and on the organelles involved in these cell death types. Then, the vulnerability of cerebellar PNs to cell death will be discussed. Then, PN death in rodent mutants, namely shaker mutant rat and lurcher mutant mouse and the therapeutic implications of the programmed cell death in PNs will be discussed.

2. Programmed cell death
2.1 Apoptosis

Apoptosis or programmed cell death type I can be triggered by various death stimuli, such as stress [9, 10], activation of death receptors [11], genetic disorders [12], and abnormally increased intracellular calcium concentration ([Ca2+]i) [13, 14, 15].

Neuronal apoptosis occurs commonly during development and maturation and appears necessary for shaping and eventual appropriate circuitry formation of the nervous system [16].

Apoptosis is characterized by distinct morphological and biochemical alterations. Morphologically, the cell shrinks, chromatin condenses, nuclear DNA fragments, cell membrane maintains its integrity, and apoptotic bodies containing nuclear material are formed. Apoptotic bodies are ultimately engulfed by phagocytes without eliciting any inflammation [17]. Biochemically, protein degradation and caspase activation are augmented characterizing apoptosis [18], which is regulated by the B-cell lymphoma 2 (Bcl-2) family proteins that comprise both anti- and pro-apoptotic proteins that should be maintained in balance, because otherwise apoptosis can be either promoted or inhibited [19, 20].

Caspases are a family of cysteine proteases, which are present in normal cells as inactive zymogens that become enzymatically active only in response to apoptotic stimuli. Active caspases are large and small fragments that result from the cleavage of the single peptide precursor, and they mediate apoptosis leading to the distinctive morphological changes [21]. According to their function, caspases can be classified as inflammatory caspases and apoptotic caspases. Inflammatory caspases are caspases-1, -4, -5, -11, -13 and -14, which mediate the proteolytic activation of inflammatory cytokines [22]. Apoptotic caspases can be either initiator or executioner caspases [21]. Apoptotic initiator caspases are caspases-2, -8, -9 and -10, that have a caspase activation and recruitment domain such as caspase -2 and -9 or a death effector domain such as caspase -8 and -10. Initiator caspases can initiate apoptosis by cleaving themselves leading to their activation and subsequent cleavage of the common downstream executioner caspases resulting in their activation [21]. Executioner caspases-3, -6 and -7, have short prodomains and lack the ability of auto-cleavage, and consequently need to be cleaved and activated by initiator caspases. Executioner caspases perform the downstream execution steps of apoptosis by cleaving multiple cellular substrates [21].

Caspases have 5 cumulative effects during apoptotic events. First, caspases can disable processes involved in homeostasis and repair, such as DNA repair processes, to prevent counterproductive events from occurring simultaneously [23]. Second, caspases can stop cell cycle progression [24]. Third, caspases can cause signal amplification and inhibitor inactivation by cleaving pro- and anti-apoptotic proteins [25]. Fourth, caspases can mediate nuclear and cytoskeletal disassembly and morphological changes [26]. Finally, caspases are involved in the recognition of dying cells for their phagocytosis and subsequent clearance [26]. Caspases cleave Ca2+-AMPA glutamate receptors, and subsequently inhibit Ca2+ build up and overload in a neuron that would lead to excitotoxicity and subsequent necrosis [27]. Caspases have also been reported to cleave and inactivate the plasma membrane Ca2+ pump (PMCA) in neurons and non-neuronal cells undergoing apoptosis [28], leading to Ca2+ build up and subsequent Ca2+ overload, which may trigger Ca2+-dependent death pathways such as necrosis, even when Ca2+ signals are not the initial death trigger. These studies have also shown that necrosis can be reduced or prevented by caspase inhibitors in brain ischemia due to caspase-mediated cleavage and inactivation of PMCAs [28]. On the other hand, inhibition of caspase activation, in the presence of cell death stimuli, can expose or even augment underlying caspase-independent death programs [29].

Caspases may also have non-apoptotic functions, such as synaptic plasticity [30, 31], and proliferation and differentiation of non-neuronal cells [32]. For instance, apoptotic-like morphological changes have been demonstrated in differentiating cells, and T-cell proliferation has been shown to be prevented by caspase inhibitors [33].

Caspase activation can be initiated by two well described apoptotic pathways (Fig. 1): the mitochondria-mediated pathway and the cell surface death receptor pathway [17]. The mitochondria-mediated apoptotic death pathway is also called the intrinsic apoptotic pathway. Permeability transition pore (PT-pore) and Bcl-2 family are important regulators of the mitochondria-mediated death pathway [34, 35]. Bcl-2 is a family of proteins that can be either pro-apoptotic, such as Bcl-2-associated X protein (Bax), or anti-apoptotic, such as Bcl-2 [36]. Members of Bcl-2 family exist on the cytoplasmic surface of different organelles, such as mitochondria, endoplasmic reticulum, and the nucleus [37]. Members of the Bcl-2 family act as regulators of the permeability transition pore (PT-pore) [38]. When the PT-pore at the mitochondrial intermembraneous contact spots is opened, small molecular weight substances, such as charged ions, are lost from the matrix causing mitochondrial depolarization and rupture of the outer mitochondrial membrane due to osmotic mitochondrial swelling [38]. The pro-apoptotic Bcl-2 family proteins regulate outer mitochondrial membrane permeabilization, causing irreversible release of apoptotic molecules, such as cytochrome c, which might be the primary regulatory step for mitochondria-mediated caspase activation [39]. Cytochrome c normally exists entirely in the intermembraneous space of mitochondria [40]. A cytosolic protein called aoptotic protease activating factor 1 (Apaf-1) binds to cytochrome c in the cytosol, in the presence of ATP [41]. Consequently, a multimeric Apaf-1/cytochrome c complex is formed, and it subsequently recruits procaspase-9, which becomes activated by proteolysis [42]. Subsequently, activated caspase-9 is released from this complex to cleave and activate executioner caspases, -3, -6 and/or -7. Cell commitment to apoptosis cannot be caused by occasional leakage of cytochrome c, due to a relatively high threshold of caspase activation set by the formation of a multimeric Apaf-1/cytochrome c complex [42].

Fig. 1.

Pathways of apoptosis. The intrinsic apoptotic pathway is triggered when apoptotic stimuli stimulate proapoptotic proteins, such as Bax, to induce the permeabilization of the outer mitochondrial membrane, and the subsequent discharge of cytochrome c from the mitochondrial intermembranous space. Cytochrome c then participates in the formation of a multimeric Apaf-1/cytochrome c complex, which then forms the apoptosome by recruiting procaspase-9 in order to activate it. Once activated, caspase-9 is dissociated from this complex, and it activates executioner caspases-3, -6, and/or -7. On the other hand, extrinsic apoptotic pathway starts by trimerization of ligand-induced receptor subsequrnt to the binding of particular death ligands to death receptors. Subsequently, particular intracellular proteins that are associated with the death receptor, such as procaspase-8, are recruited. After its recruitment, procaspase-8 is immediately activated through proteolysis. Then the active caspase-9 and active caspase-8 cleave and activate downstream executioner caspases-3, 6, and/or 7 that carry out the proteolytic events of cellular proteins and structures ultimately leading to apoptosis.

The cell surface death receptor-mediated apoptotic pathway is also called the extrinsic apoptotic pathway. Cell surface death receptors are transmembrane proteins that form a family belonging to the tumor necrosis factor (TNF)/nerve growth factor (NGF) receptor superfamily. These death receptors are activated by structurally linked ligands that belong to the TNF gene superfamily [43, 44]. Cell surface death receptor-mediated caspase activation starts by trimerization of ligand-induced receptor [43, 44]. As a consequence of ligand-death receptor binding, particular intracellular proteins that are associated with the death receptor, such as procaspase-8, are recruited. After its recruitment, procaspase-8 is immediately activated through proteolysis. Then the active form of caspase-8 cleaves and activates downstream executioner caspases [43, 44, 45].

Apoptosis is morphologically characterized by chromatin condensation, DNA fragmentation, nuclear breakup [4, 46], membrane blebbing [47], cell shrinkage [46, 48], and formation of apoptotic bodies [49]. Apoptotic bodies are membrane-bound fragments of the dying cells that are eventually cleared by phagocytosis without any associated inflammation [50].

The principal molecular components of apoptosis in neurons are the same as those in other non-neuronal cell types [51]. And the occurrence of apoptosis has been reported in many neurodegenerative diseases [51].

Autophagy and necrosis are other cell death types that have been suggested to contribute to the pathogenesis of neurodegenerative diseases in addition to apoptosis.

2.2 Autophagy

Autophagy, also called type II programmed cell death, is self-destruction of the cell [52]. Autophagy occurs normally at low basal levels as the degradative mechanism for removal of long-lived proteins and organelles [53]. Autophagy occurs in neurons in physiological processes, such as regulation of metabolism through removal of particular enzymes, morphogenesis, and tissue remodeling due to differentiation [54, 55]. Unlike apoptosis, autophagy can promote either survival or death of the cell [55]. Cells that undergo excessive autophagy, where short-lived organelles and proteins are degraded, are caused to die in a non-apoptotic manner, which is independent of caspases [55].

Based on the mechanism of autophagic vacuole formation and material transport to lysosomes, autophagy can be classified into 3 types: macroautophagy, chaperon-mediated autophagy (CMA), and microautophagy [56].

Macroautophagy is the classic autophagy, which is stimulated by stress conditions, such as nutrient deprivation, infections and toxins [57]. Intracellular stress situations, including aggregation of misfolded proteins and accumulation of damaged organelles can also induce macroautophagy [58]. Macroautophagy is characterized morphologically by the formation of double membrane-bounded vacuoles, called autophagosomes [59]. The origin of the autophagosome membrane is unknown, but it is suggested to derive from the endoplasmic reticulum, trans Golgi, and specialized membrane cisternae, called phagophores [60].

Chaperon-mediated autophagy does not require vesicle formation [61]. It is selective for a specific group of cytosolic proteins containing a pentapeptide motif [62]. Chaperon heat shock cognate 70 (Hsc70) transfers protein substrates, after recognizing that motif, to the lysosomal membrane. After binding to the lysosome-associated membrane protein 2a (lamp2a), Hsc70 translocates those proteins into the lysosomal lumen, where they are degraded by lysosomal hydrolases [62, 63, 64].

Microautophagy is the direct uptake of soluble proteins and organelles by lysosomes without autophagosome formation. The lysosomal membrane itself invaginates to engulf and subsequently degrade the cytoplasmic component [65]. These lysosomal invaginations can be observed in the electron microscope [66]. The exact physiological function of microautophagy is still to be elucidated. However, microautophagy has been suggested to play a prominent role in maintaining membrane homeostasis, by removing the outer autophagosome membrane following autophagosome-lysosome fusion in macroautophagy [65, 66].

Macroautophagy proceeds through 4 stages, which are: induction and cargo selection, vesicle formation, vesicle docking and fusion with lysosomes, and finally vesicle breakdown and recycling [56]. For the induction stage, a serine/threonine protein kinase (mTOR) acts as an important regulator of autophagy [67]. Under nutrition-rich conditions, phosphorylated mTOR negatively controls autophagy, primarily by acting on the signaling cascade that controls general translation and transicription. Under starvation conditions, mTOR is dephosphorylated and consequently inactivated leading to disinhibition of autophagy [68, 69]. Another major upstream signaling pathway that regulates starvation-induced autophagy is phosphoinositide 3-kinase (PI3K) [70]. Beclin 1 exists with class III PI3K in a complex producing one product, phosphatidylinositol 3-phosphate (PtdIns(3)P) [71]. Autophagy is stimulated by the administration of PtdIns(3)P, while it is inhibited by inhibition of class III PI3K [72, 73]. Once induced, the autophagic process begins with cargo selection. Cargo selection is the selection of the cytoplasmic part of the cell to be degraded by the autophagic process [56].

In the vesicle formation stage, a membrane sac isolation membrane or phagophore wraps around the degradation substrates and ultimately fuses to become an autophagosome or autophagic vacuole [74]. Most of the proteins involved in vesicle expansion and maturation dissociate from the mature autophagosome. Microtubule-associated protein light chain 3 (LC3) recruited on the surface of autophagosomes remains on the autophagosomal membrane. It has been shown that 2 ubiquitin-like conjugation systems are required for autophagosome formation [75, 76]. Once the autophagosome is formed, its outer membrane completely fuses with the outer lysosomal membrane making a path for the autophagic body, which is the inner membrane-bound autophagic vacuole [77]. Thus, the autophagic body is released into the lysosomal cavity [78]. Then, contents of the autophagosome are degraded by the hydrolytic enzymes in the acidic lysosomal environment. Eventually, essential cytoplasmic contents are recycled [77, 79].

Autophagy can be most reliably detected using the transmission electron microscope that illustrates the autophagic vacuoles and multilamellar whorls as the principal ultrastructural features of autophagy [80]. Multilamellar whorls form as a consequence of the compaction of engulfed organelle membranes, which typically occurs at the autolysosomal stage [81].

Disaggregated protein accumulation due to autophagic dysfunction contributes to the pathogenesis of neurodegenerative diseases such as Parkinson disease, Alzheimer disease, Huntington disease and Niemann-Pick type C disease [54, 82, 83, 84]. Niemann-Pick type C disease is a lysosomal storage disease, produced by nPN1 or nPN2 gene mutations, where neurons display characteristics of autophagic cell death [84]. Initiation of autophagy, including the induction of the transcriptional level of the autophagic genes, may use apoptotic mechanisms for the execution of cell death. For example, autophagy related 5 (Atg5) is a protein, which is the main regulator of the sequestration stage in autophagy, seems to be an important mediator of apoptosis [85, 86]. Atg5 triggers cytochrome c discharge from the mitochondrial intermembraneous space promoting mitochondria-mediated apoptosis [87]. The anti-apoptotic protein Bcl-2 inhibits Beclin-1-mediated autophagic cell death by binding to Beclin1/Atg6 [88]. On the other hand, caspase inhibition prevented cell death but did not affect vacuole formation [89].

2.3 Necrosis

Necrosis was initially called “non-lysosomal vesiculate”, since it is characterized by the destruction of the cytoplasm by non-lysosomal degradation [90]. Recently, necrosis has been recognized as a third type of programmed cell death [90]. Necroptosis is the most deeply characterized form of necrotic programmed cell death, which was first recognized as a cell death backup to apoptosis when apoptosis was inhibited by endogenous or exogenous factors. It is induced by interleukin-1β (IL-1β), TNF, particular viral infections and other factors [91]. Additionally, necroptosis is mediated by receptor-interacting serine/threonine kinase RIPK 1 to RIPK3, which assemble subsequent to the engagement of the death receptor pathway in the absence of caspase 8 forming an RIPK1/RIPK3 complex, which is called the Necrosome that is essential for activating the necroptotic pathway [92] (Fig. 2, Ref. [93, 94]). Subsequently, RIPK1 and RIPK3 undergo auto- and trans-phosphorylation, and thus permitting recruitment and activation of the pseudokinase, and eventually leading to the disruption of plasma membrane integrity, and its subsequent rupture [93]. Accordingly, endogenous molecules, which are referred to as damage-associated molecular patterns (DAMPs), are relaesed [94]. DAMPs, which include IL-1 family members, activate inflammasome, trigger inflammation and elicit immune responses [94].

Fig. 2.

Sequence of events in necroptosis. Necroptosis is mediated by receptor-interacting serine/threonine kinase RIPK 1 to RIPK3, which assemble subsequent to the engagement of the death receptor pathway in the absence of caspase 8 forming an RIPK1/RIPK3 complex, which is called the Necrosome that is essential for activating the necroptotic pathway that eventually leads to the disruption of plasma membrane integrity, and its subsequent rupture. Accordingly, endogenous molecules, which are referred to as damage-associated molecular patterns (DAMPs), are released [93, 94]. DAMPs, which include IL-1 family members, activate inflammasome, trigger inflammation and elicit immune responses.

Necroptosis has been shown to mediate a variety of human diseases, such as ischemic brain injury, systemic inflammation, ischemic reperfusion injury and neurodegenerative diseases [95].

Morphological features of necrosis include disruption of the plasma membrane early during the death process, resulting in leakage of cellular contents and subsequent inflammation, which involves microglia activation acquiring increased motility and phagocytic activity [96]. Other morphological features include enlargement of cytoplasmic organelles and detachment of ribosomes from the endoplasmic reticulum and fragmentation of the nuclear membrane [96].

2.4 Organelles involved in programmed cell death

In order to understand the interaction between the various types of cell death, this section reviews the organelles participating in different types of programmed cell death. Mitochondria, lysosomes, and endoplasmic reticulum play a significant role in the discharge and activation of death factors, as summarized in Table 1 (Ref. [77, 78, 79, 97, 98, 99, 100, 101, 102, 103, 104, 105, 106, 107, 108]). Death factors are factors signaling and mediating death pathways such as cathepsins, calpains, and other proteases.

Table 1.Roles of mitochondria, endoplasmic reticulum and lysosomes in apoptosis and autophagy.
Mitochodria References Endoplasmic reticulum References Lysosomes References
Apoptosis Increased permeabilization of the outer mitochondrial membrane triggers the release of toxic proteins from the mitochondrial intermembrane space, such as cytochrome c [97, 98] cytosolic Ca2+ overload or Ca2+ depletion from the ER lumen can stress ER function, leading to caspase-12 activation and mitochondrial membrane permeabilization [102, 103, 104] Partial and selective permeabilization of the lysosomal membrane leading to release of lysosomal proteases that can cause cell death directly by cleaving and activating caspases or indirectly by eliciting mitochondrial dysfunction and successive discharge of mitochondrial proteins [104, 107]
Autophagy Serve as membrane source of autophagosomes in starvation-induced autophagy [100] Endoplasmic reticulum–Golgi intermediate compartment (ERGIC) can contribute me- [106] Lysosomes fuse with autophagosomes, so that the the autophagic body is released into the lysosomal cavity [77, 78, 79]
Depolarization of mitochondria can induce autophagy [101] mbrane to the forming phagophore Then, contents of the autophagosome are degraded by the hydrolytic enzymes in the acidic lysosomal environment
Necroptosis Mitochondria-derived reactive oxygen species (ROS) reinforces the stability of the necrosome [99] ER stress can induce necroptosis [105] Lysosomal release of cathepsin D into cytoplasm induces necroptosis, probably by cleaving caspase-8 [108]
2.4.1 Mitochondria

Increased permeability (permeabilization) of the outer mitochondrial membrane triggers the release of toxic proteins from the mitochondrial intermembrane space, such as cytochrome c [97]. Permeabilization of the outer mitochondrial membrane can occur by 2 mechanisms. Firstly, the outer mitochondrial membrane can rupture due to opening of a permeability transition pore at the communication areas between the inner and outer mitochondrial membranes [98]. Secondly, permeabilization of the outer mitochondrial membrane can be caused by activation of proapoptotic Bcl-2 proteins, including Bax and Bak. Proapoptotic Bcl-2 proteins are normally controlled by anti-apoptotic Bcl-2 proteins, which can prevent the permeabilization of the outer mitochondrial membrane by heterodimerization with Bax-like proteins [109].

Cytochrome c is a vital constituent of the respiratory chain that normally exists in the mitochondrial intermembranous space [110]. The inhibitor of apoptosis protein family (IAP), which inhibits apoptotic pathways, controls the catalytic function of cytochrome c [111]. When released into the cytosol, cytochrome c participates in the formation of apoptosome which triggers the apoptotic cascade as reviewed earlier [112].

Apoptosis-inducing factor (AIF) normally performs oxidoreductase function in the mitochondrial intermembranous space [113]. When released in the cytosol, AIF performs apoptotic functions, which can be dependent or independent of caspases [114]. Apoptosis-inducing factor release from the intermembranous space can be triggered by lysosomal protease cathepsin D [115]. Once outside the intermembranous space, cytosolic AIF translocates to the nuclei and stimulates chromatin condensation independently of caspase cascade [116]. On the other hand, AIF release can be triggered downstream of cytochrome c due to particular proapoptotic stimuli, probably dependent on caspase cascade [117]. Additionally, Mitochondria-derived reactive oxygen species (ROS) reinforces the stability of the necrosome [99]. Additionally, mitochondria serve as membrane source of autophagosomes in starvation-induced autophagy [100]. Furthermore, depolarization of mitochondria can induce autophagy [101].

2.4.2 Endoplasmic reticulum

The endoplasmic reticulum (ER), which is the stress sensor of the cell, performs an important function in alleviating cellular stress, due to its ability to synthesize proteins [102]. Misfolded and aggregated proteins, cytosolic Ca2+ overload or Ca2+ depletion from the ER lumen can stress ER function [102]. Endoplasmic reticulum stress can initiate programmed cell death by selective activation of caspase-12, which is normally present at the cytosolic face of the ER in its inactive state [103]. Members of Bcl-2 family are present on mitochondria and ER and thereby can mediate an interaction between the two organelles [118]. As a result of this interaction, ER stress can induce mitochondrial membrane permeabilization, which is the rate-limiting step in the mitochondria-mediated apoptosis [104]. ER stress can also induce necroptosis [105]. Additionally, endoplasmic reticulum–Golgi intermediate compartment (ERGIC) can contribute membrane to the forming phagophore [106].

2.4.3 Lysosomes

The last organelles involved in programmed cell death are lysosomes containing different proteases, such as cathepsin B, L, and D [119]. Cathepsin B and D are most steady at physiological cytoplasmic pH, and appear to have a noticeable role in apoptotic and necrotic programmed cell death [119]. Oxidative stress, TNF-α and chemotherapeutic drugs can activate lysosomal proteases leading to cell death [120]. Lysosomal proteases trigger programmed cell death via multiple pathways [121]. Partial and selective permeabilization of the lysosomal membrane triggers apoptotic-like programmed cell death, whereas massive breakdown of lysosomes induces unregulated necrosis [122]. Lysosomal proteases can cause cell death directly by cleaving and activating caspases [107]. Lysosomal proteases can also endorse cell death indirectly by eliciting mitochondrial dysfunction and successive discharge of mitochondrial proteins [104]. The discharge of lysosomal cathepsins and consequent cell death can be induced by activated calpains [123]. Calpains, which are a family of Ca2+-activated neutral cytosolic proteases that normally reside in the cytosol as inactive zymogens, have been reported to perform downstream of caspase activation [124]. Furthermore, lysosomal release of cathepsin D into cytoplasm can stimulate necroptosis, probably by its contribution to the cleavage of caspase-8 [108].

And as mentioned formerly, lysosomes fuse with autophagosomes, so that the the autophagic body is released into the lysosomal cavity [77, 78]. Then, contents of the autophagosome are degraded by the hydrolytic enzymes in the acidic lysosomal environment [79].

In summary, the literature reviewed above strongly suggests that organelles involved in programmed cell death interact with each other. As a consequence, one can hypothesize that more than one type of programmed cell death may occur in the same dying cell.

3. Cerebellar PNs are among the most vulnerable neurons to cell death

Purkinje neurons are the sole output neurons of the cerebellar cortex, and they are involved in coordination of voluntary movements [125]. Death of PNs results in ataxia [126]. Purkinje neurons are the most highly susceptible population of neurons to intrinsic hereditary diseases [127] or extrinsic toxic [128], hypoxic [129], ischemic [130], and traumatic injury [131]. In addition, numerous genetic mutations directly affect PN survival and cause PN degeneration [126].

The role that Ca2+ plays in PN function and the afferents that project to PNs might be the reason why PNs are among the most vulnerable neurons to cell death [132]. Purkinje neurons receive excitatory input from two afferent fiber types, namely climbing fibers and parallel fibers [133]. Each climbing fiber makes hundreds of synaptic connections on the proximal dendrites of a single PN, resulting in a very powerful excitatory input and subsequent dramatic increase in [Ca2+]i [134]. On the other hand, parallel fibers arise from granule cells within the cerebellum, which receive the excitatory mossy fibers from different parts of the central nervous system [135]. Each PN receives input from about 175,000 parallel fibers synapsing on the distal dendritic branches of that PN [136].

Intracellular Ca2+ concentrations can be divided between the cytosolic compartment and Ca2+ storage sites such as the endoplasmic reticulum (ER) [137]. Resting [Ca2+]i is low [138]. However, there is very rapid increase in PN [Ca2+]i due to its strong stimulation by the excitatory inputs of parallel and climbing fibers afferents resulting in complex patterns of Ca2+ dynamics in PN dendrites, in which metabotropic glutamate receptor signaling pathway mediates Ca2+ influx through voltage-gated Ca2+ channels and Ca2+ release from intracellular stores [139, 140]. On the other hand, Ca2+ ions are pumped out of the cytosol by plasma membrane calcium ATPase2 (PMCA2), which is abundantly expressed in PN compared to other neurons [141]. PMCA2 uses energy in order to maintain a relatively low intracellular net Ca2+ load [141]. Reduced calcium clearance perturbs calcium dynamics in PN dendrites and disrupts the relay of information from PN to other neurons leading to neural dysfunction [141]. Indeed, partial reductions in PMCA2 levels have been shown to perturb the function and integrity of PNs and the overall outcome on PN survival [142].

Furthermore, Ca2+-binding proteins, which buffer the Ca2+ in the cell by binding to it, also participate in maintaining [Ca2+]i that are essential for normal Ca2+ signaling, and thus preventing the accumulation of pathologic quantities of Ca2+ [138, 143, 144, 145].

Although PNs are highly susceptible to cell death, not all PNs respond similarly to death stimuli. For example, only specific PNs that are zebrin 2- and excitatory amino-acid transporter 4 (EAAT4)-negative are particularly sensitive to hypoxic stress caused by cerebral ischemia [146]. Additionally, while some PNs die in response to a death stimulus; other PNs survive indicating the presence of heterogeneous PN populations.

4. PN death in neurodegenerative diseases in animal mutants

Mechanisms underlying PN neurodegeneration is better investigated in animal model studies, because autopsy delay of post-mortem cerebella leads to artifacts and shows end-stage disease only. Neurodegenerative diseases affect the cerebellum in many species, in which many single mutations totally affect different genes. These animal mutants have demonstrated distinctive age of onset of PN degeneration, in addition to distinctive temporal and spatial patterns of PN degeneration. Such animal models have been used to illustrate the type(s) of programmed cell death causing PN loss and the pathogenesis of the particular neurodegenerative disease (Table 2, Ref. [147, 148, 149, 150, 151, 152]). Additionally, those animal models allow targeted studies of different pathological mechanisms underlying PN death. Thus, PN death in lurcher mutant mice and shaker mutant rats, as examples of animal mutations that lead to PN death, will be reviewed in this section.

Table 2.Experimental evidence for the occurrence of apoptosis and autophagy in lurcher mutant rat and shaker mutant rat.
Lurcher mutant mouse Reference Shaker mutant rat Reference
Apoptosis Morphological features of apoptosis in the electron micrographs of degenerating PNs [147] Morphological features of apoptosis in the electron micrographs of degenerating PNs [151, 152]
TUNEL labeling in PN [148] TUNEL labeling in PN [152]
Active caspase-3 expression in PN [148] Active caspase-3 expression in PN [152]
Increased Bax expression in PN [149]
Autophagy Autophagic macromolecular complexes were formed by interactions between GluRδ2 and Beclin1 via nPIST [150] Morphological features of autophagy in the electron micrographs of degenerating PNs [151]
Morphological features of autophagy in the electron micrographs of degenerating PNs [150]
4.1 Lurcher mutant mouse

Lurcher is an autosomal semidominant mutation. Homozygous lurcher mice die shortly after birth due to a massive loss of mid- and hindbrain neurons during late embryogenesis. In heterozygous lurcher mutant mice, starting during the second postnatal week, PNs degenerate in a rapid and extensive manner, so that 90% of PNs have already degenerated by 26 postnatal days of age [153]. No PNs survive at 90 postnatal days of age [154]. PNs have been suggested to die by an excitotoxic mechanism, since the lurcher mutation is a gain of function mutation that transforms the glutamate receptor delta2 (GluRδ2) subunit into a constitutively open channel that maintains PNs at a depolarized resting membrane potential [153]. In addition to the cell autonomous PN death, target-related cell death occurs in 90% of granule cells and 75% of inferior olivary neurons in the lurcher mutant mouse [155]. Both apoptosis and autophagy have been reported in lurcher mutant PNs. The GluRδ2Lc mutation, naturally occurring postnatal cell death, and the physical trauma of making organotypic slice cultures have been shown to induce the multiple PN death pathways [156].

Morphological abnormalities indicative of apoptosis including the appearance of perinuclear clumps of chromatin, axonal swellings (torpedoes) and a delayed maturation process, have been observed in lurcher PNs at 8 postnatal days of age [147]. Delayed maturation process of the lurcher mutant PNs was revealed by the incomplete development of the basal polysomal mass of PN bodies and the late development of proximal and distal compartments of the dendritic trees, which were hyperspinous [157]. Swollen mitochondria with dilated cristae, reported in some models of apoptosis [158], were also observed in dying lurcher PNs [157]. Norman et al. [159] demonstrated markedly condensed nuclear chromatin, and nuclear and cell membrane blebbing in dying lurcher PNs at 12 postnatal days of age. The presence of deoxyribonucleic acid (DNA) fragmentation, detected by TUNEL, active caspase-3 expression, and increased Bax expression have been reported in lurcher mice, suggesting that PNs die by apoptosis [148, 149].

Autophagy was also reported in lurcher PNs when autophagic macromolecular complexes were observed to be formed by interactions between GluRδ2 and Beclin1 via the neuronal isoform of protein-interacting specifically with TC10 (nPIST) [150]. Beclin1 is the human ortholog of the yeast autophagic gene Atg6 [160]. Autophagy was introduced in cells transfected with the lurcher mutant glutamate receptor (GluRδ2Lc) but not wild type GluRδ2, and that the death of those transfected cells could be partially rescued pharmacologically by inhibiting autophagy with 3-methyladenine. The occurrence of moderate autophagy has, indeed, been suggested by demonstrating autophagosomes in the electron micrographs of degenerating PNs in the lurcher mutant mouse [150].

Furthermore, an attempt to rescue lurcher PNs from death was by the genetic elimination of tissue plasminogen activator (tPA) [161], which is a serine protease member of the fibrinolytic system that mediates excitotoxic neuronal cell death. Lu and Tsirka (2002) crossed lurcher mutant mice with tPA-null mutant mice to eliminate tPA in lurcher mutant mice (Lc/+; tPA–/–) [161]. Lurcher homozygotes lacking tPA (Lc/Lc; tPA–/–) died shortly after birth. At 12 and 30 postnatal days of age, significantly lower numbers of surviving PNs in both lurcher mutant mice (Lc/+; tPA+/+) and tPA-null lurcher double mutants (Lc/+; tPA–/–) compared with the wild-type mice were reported [161]. However, PN death was clearly attenuated in tPA-null lurcher double mutants (Lc/+; tPA–/–) compared with the lurcher mutants (Lc/+; tPA+/+) [161]. Additionally, the tissue plasminogen activator elimination resulted in a decrease in active caspase-8 expression but had no effect on active caspase-9 expression at 12 and 30 postnatal days of age [161].

Moreover, previous studies critically tested the role of various cell death pathways in lurcher mutant PN degeneration with respect to evidence for the molecular heterogeneity of Purkinje cells [162]. The expression of putative survival factors, such as heat shock proteins, in a subset of cerebellar PNs was proposed to influence cell death pathways and attribute to the pattern and diverse mechanisms of lurcher mutant PN degeneration [162].

4.2 Shaker mutant rat

Shaker mutant rats have an x-linked recessive mutation [163], in which missense change is identified in ATPase plasma membrane Ca2+ transporting 3 (Atp2b3) gene, which encodes PMCA3 that is highly expressed in the cerebellum [164]. However, the resultant PMCA3 variant has been shown in the in vitro analysis not to have any functional effects, suggesting that this genetic change probably remains not causative, but very closely associated with the causative shaker mutation [165]. The shaker mutation naturally results in the degeneration of spatially limited populations of cerebellar PNs at an interval of seven to fourteen weeks of postnatal age [163]. Nevertheless, PN degeneration was reported to occur at earlier [166] or later ages [167] in response to experimental conditions. Purkinje neuron degeneration results in gait ataxia and whole body tremor, which develop concomitantly with PN degeneration [163]. At risk PNs always degenerate and are located in two cortical areas that are known as anterior (ADC) and posterior (PDC) degeneration compartments [163]. The ADC comprises lobules I-VIb and d, whereas the PDC comprises lobules VIIb-VIII and IXa-c. At risk PNs in the PDC normally degenerate 1–2 weeks later and slower compared to the ADC at risk PNs [163]. On the other hand, secure PNs usually survive, and they exist outside the ADC and PDC forming an intermediate (ISC) and a flocculonodular (FNSC) survival compartments [163]. The ISC comprises lobules VIc-VIIa, while the FNSC comprises lobules X and IXd [163]. Shaker mutant PNs have been shown to die by multiple death pathways including apoptosis and autophagy [151, 152].

Previous studies sought to identify epigenetic factors that may modify the events leading to hereditary PN degeneration in the shaker mutant rat. These studies included chronic infusion of neurotrophic factors [167] and inferior olive chemoablation [166].

Trophic factors play critical roles for neuronal survival developmentally, post-developmentally, and therapeutically [168]. Trophic factors were used in an attempt to protect at risk PNs against cell death in the shaker mutant rat [167].

Exogenous glial derived neurotrophic factor (GDNF) and/or insulin-like growth factor 1 (IGF-1) were chronically infused intraventricularly by osmotic pumps to the 5 week 3 day old shaker mutant rat cerebellum [167]. Four weeks of continued infusion of GDNF or IGF-1 delayed the degeneration of many at risk PNs in the ADC [167]. Unexpectedly, the number and spatial distribution of surviving PNs were reported [167] to be comparable to that found in age-matched non-saline and saline-infused controls 2 weeks and 4 weeks after stopping GDNF or IGF-I infusion. Eight weeks of continued trophic factor infusion did not support the continued survival of most PNs in the ADC (165). In addition, four weeks of continued infusion of both GDNF and IGF-1 delayed the degeneration of many more PNs than with either neurotrophic factor alone, suggesting that GDNF and IGF-1 acted on disparate mutant PN populations, whose differential survival influences various aspects of locomotion [167]. Thus, trophic factors were proposed to be only partially and transiently neuroprotective for at risk PNs.

Climbing fibers, which originate from the inferior olivary neurons of medulla oblongata, dynamically control PN structure and function [169]. Climbing fiber deafferentation resulting from IO chemoablation causes temporal acceleration of the hereditary degeneration of at risk PNs in shaker mutant rats [166]. Temporally accelerated PN death caused by IO chemoablation in shaker mutant rats could not be blocked by chronic infusion of GDNF and IGF-1 prior to IO chemoablation suggesting that olivocerebellar deafferentation might have led to an overexcitation of parallel fibers due to the removal of inhibition by the climbing fibers [170].

Morphological evidence was provided for the occurrence of apoptosis and autophagy in at risk PNs during the natural phenotypic expression of the shaker mutation [151]. Additionally, active caspase-3 immunoreactivity and DNA fragmentation, which was detected by TUNEL labeling, were shown in at risk PNs, suggesting an evidence for active caspase-3-mediated apoptosis in at risk PNs during the ordinary phenotypic expression of the shaker mutation [152]. Furthermore, active caspase-3 expression was augmented in at risk PNs following IO chemoablation [171].

5. Therapeutic implications

Since the end-point of apoptosis, autophagy, and/or necrosis in PN is their death, therapeutic approaches that target the molecular and biochemical events of the ongoing one or more type of programmed cell death in PNs may protect against their loss.

Experiments have attempted to genetically rescue lurcher mutant PNs from death using transgenic mice overexpressing anti-apoptotic Bcl-2 protein [172], lacking pro-apoptotic Bax protein [173], or lacking tissue plasminogen activator (tPA) (Table 3, Ref. [161, 166, 167, 170, 172, 173, 174, 175]) [161]. Zanjani et al. [172] bred lurcher mutant mice with human Bcl-2 transgenic mice, to produce lurcher mutant mice overexpressing Bcl-2 (+/Lc-Hu-Bcl-2), which is a key regulator of apoptosis. Zanjani et al. [172] used lurcher (+/Lc) mutant mice as controls. At 2 postnatal months of age, substantial numbers of PNs were reported to survive in the lurcher mutant overexpressing human Bcl-2 (+/Lc-Hu-Bcl-2), whereas virtually all PNs had degenerated in the control (+/Lc). Surviving PNs overexpressing Bcl-2 were distributed throughout all lobules and the different parasagittal planes of the cerebellum. However, by 5–6 postnatal months of age, only very few surviving PNs were reported and the number of surviving PNs in +/Lc-Hu-Bcl-2 was comparable to that observed in controls (+/Lc). These findings [172] suggested that Bcl-2 overexpression could temporarily delay PN degeneration, and Bcl-2 overexpression was not adequate to permanently protect lurcher mutant PNs from the effects of the leaky GluRδ2 receptor.

Table 3.Potential implications in the lurcher mutant rat and shaker mutant rat.
Therapeutic implications Reference
Lurcher mutant mouse Bcl-2 overexpression could temporarily delay lurcher PN degeneration [172]
Bax deletion only transiently delayed the degeneration of lurcher PNs [173]
tPA elimination could delay death receptor-mediated apoptotic pathway and to ultimately trigger other apoptotic pathways that were sufficient to elicit PN death [161]
Autophagic cell death in granule cells in vitro served as an alternate death pathway, when apoptosis was blocked by caspase inhibition [174]
3-methyladenine inhibition of autophagy in HeLa cell culture triggered apoptotic cell death [175]
Shaker mutant rat Trophic factors (GDNF and IGF-1) could be only partially and transiently neuroprotective for at risk PNs [167]
Climbing fiber deafferentation caused temporal acceleration of the hereditary degeneration of at risk PNs [166]
Temporally accelerated PN death caused by IO chemoablation in shaker mutant rats could not be blocked by chronic infusion of GDNF and IGF-1 prior to IO chemoablation [170]

A different attempt to rescue lurcher PNs from death/degeneration used heterozygous lurcher mutant mice (+/Lc) bred with Bax knockout mice (Bax–/–) to generate Bax-knockout lurcher double mutant mice (+/Lc, Bax–/–) [173]. Bax deletion was reported to be insufficient to inhibit the death of brainstem neurons in lurcher homozygotes, which were never found in the litters. At 15 postnatal days of age, there were significantly more surviving PNs in +/Lc, Bax–/– than in control lurcher mutants. The increased number of surviving PNs was throughout all regions of the cerebellum in double mutants (+/Lc, Bax–/–), indicating that Bax deletion did not rescue particular subpopulation of PNs. By 30 postnatal days of age, the number of surviving PNs in the double mutant mice decreased to a low level, which was not significantly different from that found in control lurcher mutant mice [173]. These findings suggested that Bax deletion did not permanently inhibit apoptosis but only transiently delayed the degeneration of lurcher PNs for several weeks [173].

Moreover, lurcher mutation was suggested to cause depolarized resting membrane potentials due to the constitutively open GluRδ2 channels in PNs [161]. This depolarization triggered the intrinsic apoptotic pathway as well as Ca2+-dependent secretion of tPA [161]. Secreted tPA initiated the extrinsic apoptotic pathway and caspase-8 activation, which facilitated the initial PN degeneration observed in lurcher mutant mice [161]. Therefore, tPA elimination was suggested to delay death receptor-mediated apoptotic pathway and to ultimately trigger other apoptotic pathways that were sufficient to elicit PN death [161].

Autophagic cell death in granule cells in vitro served as an alternate death pathway, when apoptosis was blocked by caspase inhibition [174]. Alternatively, 3-methyladenine inhibition of autophagy in HeLa cell culture triggered apoptotic cell death [175]. Moreover, a study on Huntington disease suggested that prothymosin-α might cause cells to shift between apoptosis and autophagy by adverse regulation of the apoptosome activity [176].

Thus, treatments that interfere with the pathways of the various types of programmed cell death may denote promising therapeutic approaches in the protection against PN loss in various neurodegenerative diseases in patients. However, it must be recognized that there are concerns about targeting programmed cell death in neurodegenerative diseases; since it is still to be comprehended whether or not inhibition of the different types of programmed cell death in PNs can be effective and harmless. Thus, a careful assessment in this regard is essential.

6. Conclusions

Purkinje neurons are among the most vulnerable neurons to cell death due to intrinsic hereditary diseases or extrinsic factors. Purkinje neuronal death causes cerebellar abnormalities, which are manifested by abnormalities in gait and posture and thus can be clearly recognized. Programmed cell death is proposed to be involved in the pathogenesis of neurodegenerative diseases associated with PN loss, based on in vitro, in vivo, and human postmortem studies [151, 152, 162, 177]. A better understanding of the molecular events regulating the various types of programmed cell death would be extremely beneficial for disease therapy, as these mechanisms are implicated in a variety of human neurodegenerative diseases, by making it possible to identify the possible factors that can be targeted therapeutically in order to halt or slow the progression of the disease. Further research in this area can be anticipated to provide promising opportunities regarding therapeutic manipulation of cell death programs.

Abbreviations

PNs, Purkinje neurons; PD, Parkinson disease; AD, Alzheimer disease; HD, Huntington disease; Bcl-2, B-cell lymphoma 2; PMCA, plasma membrane Ca2+ pump; PT-pore, Permeability transition pore; Apaf-1, apoptotic protease activating factor 1; TNF, tumor necrosis factor; NGF, nerve growth factor; Hsc70, heat shock cognate 70; lamp2a, lysosome-associated membrane protein 2a; PI3K, phosphoinositide 3-kinase; PtdIns(3)P, phosphatidylinositol 3-phosphate; Atg5, autophagy related 5; RIPK, receptor-interacting serine/threonine kinase; DAMPs, damage-associated molecular patterns; ER, endoplasmic reticulum; ERGIC, endoplasmic reticulum–Golgi intermediate compartment; [Ca2+]i, intracellular calcium concentration; EAAT4, excitatory amino-acid transporter 4; GluRδ2, glutamate receptor delta2; DNA, deoxyribonucleic acid; nPIST, neuronal isoform of protein-interacting specifically with TC10; GluRδ2Lc, lurcher mutant glutamate receptor; tPA, tissue plasminogen activator; Lc/Lc; tPA–/–, Lurcher homozygotes lacking tPA; Lc/+; tPA+/+, lurcher mutant mice; Lc/+; tPA–/–, tPA-null lurcher double mutants; Atp2b, ATPase plasma membrane Ca2+ transporting 3; ADC, anterior degeneration compartment; PDC, posterior degeneration compartment; ISC, intermediate survival compartment; FNSC, flocculonodular survival compartment; GDNF, glial derived neurotrophic factor; IGF-1, insulin-like growth factor 1; +/Lc-Hu-Bcl-2, lurcher mutant overexpressing human Bcl-2; +/Lc, heterozygous lurcher mutant mice; Bax–/–, Bax knockout mice; +/Lc, Bax–/–, Bax-knockout lurcher double mutant mice.

Ethics approval and consent to participate

Not applicable.

Acknowledgment

We thank the five anonymous reviewers for the excellent criticism of the article.

Funding

This research received no external funding.

Conflict of interest

The author declares no conflict of interest.

References
[1]
Dugger BN, Dickson DW. Pathology of neurodegenerative diseases. Cold Spring Harbor Perspectives in Biology. 2017; 9: a028035.
[2]
Sullivan R, Yau WY, O’Connor E, Houlden H: spinocerebellar ataxia: an update. Journal of Neurology. 2019; 266: 533–544.
[3]
Kumar A, Dhawan A, Kadam A, Shinde A: Autophagy and mitochondria: targets in eurodegenerative disorders. CNS & Neurological Disorders Drug Targets. 2018; 17: 696–705.
[4]
D’Arcy MS: Cell death: a review of the major forms of apoptosis, necrosis and autophagy. Cell Biology International. 2019; 43: 582–592.
[5]
Puyal J, Ginet V, Grishchuk Y, Truttmann AC, Clarke PG: Neuronal autophagy as a mediator of life and death: contrasting roles in chronic neurodegenerative and acute neural disorders. Neuroscientist. 2012; 18: 224–236.
[6]
Xing S, Zhang Y, Li J, Zhang J, Li Y, Dang C, et al. Beclin 1 knockdown inhibits autophagic activation and prevents the secondary neurodegenerative damage in the ipsilateral thalamus following focal cerebral infarction. Autophagy. 2012; 8: 63–76.
[7]
Ou L, Lin S, Song B, Liu J, Lai R, Shao L. The mechanisms of graphene-based materials-induced programmed cell death: a review of apoptosis, autophagy, and programmed necrosis. International Journal of Nanomedicine. 2017; 12: 6633–6646.
[8]
Galluzzi L, Vitale I, Abrams JM, Alnemri ES, Baehrecke EH, Blagosklonny MV, et al. Molecular definitions of cell death subroutines: recommendations of the Nomenclature Committee on Cell Death 2012. Cell Death and Differentiation. 2012; 19: 107–120.
[9]
Khuzhakhmetova LK, Teply DL, Bazhanova ED. Pharmacological correction of alterations of apoptosis of the neurons of the hypothalamus suprachiasmatic nucleus and pinealocytes during aging and stress. Advances in Gerontology. 2019; 32: 915–922. (In Russian)
[10]
Liu W, Yang T, Xu Z, Xu B, Deng Y. Methyl-mercury induces apoptosis through ROS-mediated endoplasmic reticulum stress and mitochondrial apoptosis pathways activation in rat cortical neurons. Free Radical Research. 2019; 53: 26–44.
[11]
Yin F, Zhou H, Fang Y, Li C, He Y, Yu L, et al. Astragaloside IV alleviates ischemia reperfusion-induced apoptosis by inhibiting the activation of key factors in death receptor pathway and mitochondrial pathway. Journal of Ethnopharmacology. 2020; 248: 112319.
[12]
Caneus J, Granic A, Rademakers R, Dickson DW, Coughlan CM, Chial HJ, et al. Mitotic defects lead to neuronal aneuploidy and apoptosis in frontotemporal lobar degeneration caused by MAPT mutations. Molecular Biology of the Cell. 2018; 29: 575–586.
[13]
Wang X, Chen J, Wang H, Yu H, Wang C, You J, et al. Memantine can Reduce Ethanol-Induced Caspase-3 Activity and Apoptosis in H4 Cells by Decreasing Intracellular Calcium. Journal of Molecular Neuroscience. 2017; 62: 402–411.
[14]
Jiang L, Zhong J, Dou X, Cheng C, Huang Z, Sun X. Effects of ApoE on intracellular calcium levels and apoptosis of neurons after mechanical injury. Neuroscience. 2015; 301: 375–383.
[15]
Fang X, Wu C, Li H, Yuan W, Wang X. Elevation of intracellular calcium and oxidative stress is involved in perfluorononanoic acid–induced neurotoxicity. Toxicology and Industrial Health. 2018; 34: 139–145.
[16]
Castagna C, Merighi A, Lossi L. Cell death and neurodegeneration in the postnatal development of cerebellar vermis in normal and Reeler mice. Annals of Anatomy - Anatomischer Anzeiger. 2016; 207: 76–90.
[17]
Erekat NS. Apoptosis and its role in Parkinson’s disease. In: Stoker TB, Greenland JC, (eds) Parkinson’s Disease: Pathogenesis and Clinical Aspects. Brisbane (AU): Codon Publications. 2018.
[18]
Redza-Dutordoir M, Averill-Bates DA. Activation of apoptosis signalling pathways by reactive oxygen species. Biochimica Et Biophysica Acta. 2016; 1863: 2977–2992.
[19]
Hengartner MO. The biochemistry of apoptosis. Nature. 2000; 407: 770–776.
[20]
Elmore S. Apoptosis: a review of programmed cell death. Toxicologic Pathology. 2007; 35: 495–516.
[21]
Julien O, Wells JA. Caspases and their substrates. Cell Death & Differentiation. 2017; 24: 1380–1389.
[22]
Nadiri A, Wolinski MK, Saleh M. The inflammatory caspases: key players in the host response to pathogenic invasion and sepsis. Journal of Immunology. 2006; 177: 4239–4245.
[23]
Kopeina GS, Prokhorova EA, Lavrik IN, Zhivotovsky B. Alterations in the nucleocytoplasmic transport in apoptosis: Caspases lead the way. Cell Proliferation. 2018; 51: e12467.
[24]
Podmirseg SR, Jäkel H, Ranches GD, Kullmann MK, Sohm B, Villunger A, et al. Caspases uncouple p27Kip1 from cell cycle regulated degradation and abolish its ability to stimulate cell migration and invasion. Oncogene. 2016; 35: 4580–4590.
[25]
Wolf BB, Green DR. Suicidal tendencies: apoptotic cell death by caspase family proteinases. Journal of Biological Chemistry. 1999; 274: 20049–20052.
[26]
Nicholson DW, Thornberry NA. Caspases: killer proteases. Trends in Biochemical Sciences. 1997; 22: 299–306.
[27]
Glazner GW, Chan SL, Lu C, Mattson MP. Caspase-mediated degradation of AMPA receptor subunits: a mechanism for preventing excitotoxic necrosis and ensuring apoptosis. Journal of Neuroscience. 2000; 20: 3641–3649.
[28]
Schwab BL, Guerini D, Didszun C, Bano D, Ferrando-May E, Fava E, et al. Cleavage of plasma membrane calcium pumps by caspases: a link between apoptosis and necrosis. Cell Death & Differentiation. 2002; 9: 818–831.
[29]
Scheller C, Knöferle J, Ullrich A, Prottengeier J, Racek T, Sopper S, et al. Caspase inhibition in apoptotic T cells triggers necrotic cell death depending on the cell type and the proapoptotic stimulus. Journal of Cellular Biochemistry. 2006; 97: 1350–1361.
[30]
Li Z, Sheng M. Caspases in synaptic plasticity. Molecular Brain. 2012; 5: 15.
[31]
Unsain N, Barker PA. New views on the misconstrued: executioner caspases and their diverse non-apoptotic roles. Neuron. 2015; 88: 461–474.
[32]
Larsen BD, Sorensen CS. The caspase-activated DNase: apoptosis and beyond. FEBS Journal. 2017; 284: 1160–1170.
[33]
Fischer U, Jänicke RU, Schulze-Osthoff K. Many cuts to ruin: a comprehensive update of caspase substrates. Cell Death & Differentiation. 2003; 10: 76–100.
[34]
Grimm S, Brdiczka D. The permeability transition pore in cell death. Apoptosis. 2007; 12: 841–855.
[35]
Forte M, Bernardi P. The permeability transition and BCL-2 family proteins in apoptosis: co-conspirators or independent agents? Cell Death & Differentiation. 2006; 13: 1287–1290.
[36]
Adams CM, Clark-Garvey S, Porcu P, Eischen CM. Targeting the Bcl-2 family in B cell lymphoma. Frontiers in Oncology. 2018; 8: 636.
[37]
Popgeorgiev N, Jabbour L, Gillet G. Subcellular localization and dynamics of the Bcl-2 family of proteins. Frontiers in Cell and Developmental Biology. 2018; 6: 13.
[38]
Martinou J, Youle R. Mitochondria in apoptosis: Bcl-2 family members and mitochondrial dynamics. Developmental Cell. 2011; 21: 92–101.
[39]
García-Sáez AJ, Fuertes G, Suckale J, Salgado J. Permeabilization of the outer mitochondrial membrane by Bcl-2 proteins. Advances in Experimental Medicine and Biology. 2010; 677: 91–105.
[40]
Andreyev A, Tamrakar P, Rosenthal RE, Fiskum G. Calcium uptake and cytochrome c release from normal and ischemic brain mitochondria. Neurochemistry International. 2018; 117: 15–22.
[41]
Yadav N, Gogada R, O’Malley J, Gundampati RK, Jayanthi S, Hashmi S, et al. Molecular insights on cytochrome c and nucleotide regulation of apoptosome function and its implication in cancer. Biochimica Et Biophysica Acta. 2020; 1867: 118573.
[42]
Rubio C, Mendoza C, Trejo C, Custodio V, Rubio-Osornio M, Hernández L, et al. Activation of the extrinsic and intrinsic apoptotic pathways in cerebellum of kindled rats. The Cerebellum. 2019; 18: 750–760.
[43]
Leonard BC, Johnson DE. Signaling by cell surface death receptors: Alterations in head and neck cancer. Advances in Biological Regulation. 2018; 67: 170–178.
[44]
Zhou X, Jiang W, Liu Z, Liu S, Liang X. Virus infection and death receptor-mediated apoptosis. Viruses. 2017; 9: 316.
[45]
Erekat NS. Cerebellar upregulation of cell surface death receptor-mediated apoptotic factors in harmaline-induced tremor: an immunohistochemistry study. Journal of Cell Death. 2018; 11: 1179066018809091.
[46]
Yadav PK, Tiwari M, Gupta A, Sharma A, Prasad S, Pandey AN, et al. Germ cell depletion from mammalian ovary: possible involvement of apoptosis and autophagy. Journal of Biomedical Science. 2018; 25: 36.
[47]
Fakai MI, Abd Malek SN, Karsani SA. Induction of apoptosis by chalepin through phosphatidylserine externalisations and DNA fragmentation in breast cancer cells (MCF7). Life Sciences. 2019; 220: 186–193.
[48]
Teoh PL, Liau M, Cheong BE. Phyla nodiflora L. Extracts induce apoptosis and cell cycle arrest in human breast cancer cell line, MCF-7. Nutrition and Cancer. 2019; 71: 668–675.
[49]
Xu X, Lai Y, Hua Z. Apoptosis and apoptotic body: disease message and therapeutic target potentials. Bioscience Reports. 2019; 39: BSR20180992.
[50]
Tsuruyama T, Okamoto S, Fujimoto Y, Yoshizawa A, Yoshitoshi E, Egawa H, et al. Histology of intestinal allografts: lymphocyte apoptosis and phagocytosis of lymphocytic apoptotic bodies are diagnostic findings of acute rejection in addition to crypt apoptosis. American Journal of Surgical Pathology. 2013; 37: 178–184.
[51]
Ghavami S, Shojaei S, Yeganeh B, Ande SR, Jangamreddy JR, Mehrpour M, et al. Autophagy and apoptosis dysfunction in neurodegenerative disorders. Progress in Neurobiology. 2014; 112: 24–49.
[52]
Chen Y, Hua Y, Li X, Arslan IM, Zhang W, Meng G. Distinct types of cell death and the implication in diabetic cardiomyopathy. Frontiers in Pharmacology. 2020; 11: 42.
[53]
Ravanan P, Srikumar IF, Talwar P. Autophagy: the spotlight for cellular stress responses. Life Sciences. 2017; 188: 53–67.
[54]
Stavoe AKH, Holzbaur ELF. Autophagy in neurons. Annual Review of Cell and Developmental Biology. 2019; 35: 477–500.
[55]
Nikoletopoulou V, Papandreou M, Tavernarakis N. Autophagy in the physiology and pathology of the central nervous system. Cell Death & Differentiation. 2015; 22: 398–407.
[56]
Ueno T, Komatsu M. Monitoring autophagy flux and activity: principles and applications. BioEssays. 2020; 42: 2000122.
[57]
Meng T, Lin S, Zhuang H, Huang H, He Z, Hu Y, et al. Recent progress in the role of autophagy in neurological diseases. Cell Stress. 2020; 3: 141–161.
[58]
Kesidou E, Lagoudaki R, Touloumi O, Poulatsidou K, Simeonidou C. Autophagy and neurodegenerative disorders. Neural Regeneration Research. 2013; 8: 2275–2283.
[59]
Kalachev AV, Yurchenko OV, Kiselev KV. Macroautophagy is involved in residual bodies formation during spermatogenesis in sea urchins, Strongylocentrotus intermedius. Tissue & Cell. 2019; 56: 79–82.
[60]
Wei Y, Liu M, Li X, Liu J, Li H. Origin of the autophagosome membrane in mammals. BioMed Research International. 2018; 2018: 1012789.
[61]
Tasset I, Cuervo AM. Role of chaperone-mediated autophagy in metabolism. The FEBS Journal. 2016; 283: 2403–2413.
[62]
Sun L, Lian Y, Ding J, Meng Y, Li C, Chen L, et al. The role of chaperone‐mediated autophagy in neurotoxicity induced by alpha‐synuclein after methamphetamine exposure. Brain and Behavior. 2019; 9: e01352.
[63]
Losmanová T, Janser FA, Humbert M, Tokarchuk I, Schläfli AM, Neppl C, et al. Chaperone-mediated autophagy markers LAMP2a and HSC70 are independent adverse prognostic markers in primary resected squamous cell carcinomas of the lung. Oxidative Medicine and Cellular Longevity. 2020; 2020: 8506572.
[64]
Fernández-Fernández MR, Gragera M, Ochoa-Ibarrola L, Quintana-Gallardo L, Valpuesta JM. Hsp70—a master regulator in protein degradation. FEBS Letters. 2017; 591: 2648–2660.
[65]
Schuck S. Microautophagy—distinct molecular mechanisms handle cargoes of many sizes. Journal of Cell Science. 2020; 133: jcs246322.
[66]
Oku M, Sakai Y. Three Distinct types of microautophagy based on membrane dynamics and molecular machineries. BioEssays. 2018; 40: e1800008.
[67]
Sridharan S, Jain K, Basu A. Regulation of autophagy by kinases. Cancers. 2011; 3: 2630–2654.
[68]
Kim YC, Guan K. MTOR: a pharmacologic target for autophagy regulation. The Journal of Clinical Investigation. 2015; 125: 25–32.
[69]
Dunlop EA, Tee AR. MTOR and autophagy: a dynamic relationship governed by nutrients and energy. Seminars in Cell & Developmental Biology. 2014; 36: 121–129.
[70]
McManus S, Bisson M, Chamberland R, Roy M, Nazari S, Roux S. Autophagy and 3-phosphoinositide-dependent kinase 1 (PDK1)-related kinome in pagetic osteoclasts. Journal of Bone and Mineral Research. 2016; 31: 1334–1343.
[71]
Birgisdottir AB, Mouilleron S, Bhujabal Z, Wirth M, Sjøttem E, Evjen G, et al. Members of the autophagy class III phosphatidylinositol 3-kinase complex i interact with GABARAP and GABARAPL1 via LIR motifs. Autophagy. 2019; 15: 1333–1355.
[72]
Yu P, Zhang Y, Li C, Li Y, Jiang S, Zhang X, et al. Class III PI3K-mediated prolonged activation of autophagy plays a critical role in the transition of cardiac hypertrophy to heart failure. Journal of Cellular and Molecular Medicine 2015; 19: 1710-1719.
[73]
Polajnar M, Žerovnik E. Impaired autophagy: a link between neurodegenerative and neuropsychiatric diseases. Journal of Cellular and Molecular Medicine. 2014; 18: 1705–1711.
[74]
Amaya C, Fader CM, Colombo MI. Autophagy and proteins involved in vesicular trafficking. FEBS Letters. 2015; 589: 3343–3353.
[75]
Nakatogawa H. Two ubiquitin-like conjugation systems that mediate membrane formation during autophagy. Essays in Biochemistry. 2013; 55: 39–50.
[76]
Kuma A, Komatsu M, Mizushima N. Autophagy-monitoring and autophagy-deficient mice. Autophagy. 2017; 13: 1619–1628.
[77]
Turco E, Martens S. Insights into autophagosome biogenesis from in vitro reconstitutions. Journal of Structural Biology. 2016; 196: 29–36.
[78]
Kudriaeva AA, Sokolov AV, Belogurov AAJ. Stochastics of degradation: the autophagic-lysosomal system of the cell. Acta Naturae. 2020; 12: 18–32.
[79]
Nakatogawa H. Mechanisms governing autophagosome biogenesis. Nature Reviews Molecular Cell Biology. 2020; 21: 439–458.
[80]
Papini A. Investigation of morphological features of autophagy during plant programmed cell death. Methods in Molecular Biology. 2019; 1743: 9–19.
[81]
Jung M, Choi H, Mun JY. The autophagy research in electron microscopy. Applied Microscopy. 2018; 49: 11.
[82]
Martinez-Vicente M. Autophagy in neurodegenerative diseases: from pathogenic dysfunction to therapeutic modulation. Seminars in Cell & Developmental Biology. 2015; 40: 115–126.
[83]
Malik BR, Maddison DC, Smith GA, Peters OM. Autophagic and endo-lysosomal dysfunction in neurodegenerative disease. Molecular Brain. 2019; 12: 100.
[84]
Sarkar S, Carroll B, Buganim Y, Maetzel D, Ng AHM, Cassady JP, et al. Impaired autophagy in the lipid-storage disorder Niemann-Pick type C1 disease. Cell Reports. 2013; 5: 1302–1315.
[85]
El-Khattouti A, Selimovic D, Haikel Y, Hassan M. Crosstalk between apoptosis and autophagy: molecular mechanisms and therapeutic strategies in cancer. Journal of Cell Death. 2013; 6: 37–55.
[86]
Wang Y, Qin Z. Coordination of autophagy with other cellular activities. Acta Pharmacologica Sinica. 2013; 34: 585–594.
[87]
Shi M, Zhang T, Sun L, Luo Y, Liu D, Xie S, et al. Calpain, Atg5 and Bak play important roles in the crosstalk between apoptosis and autophagy induced by influx of extracellular calcium. Apoptosis. 2013; 18: 435–451.
[88]
Zhang J, Zhang S, Shi Q, Allen TD, You F, Yang D. The anti-apoptotic proteins Bcl-2 and Bcl-xL suppress Beclin 1/Atg6-mediated lethal autophagy in polyploid cells. Experimental Cell Research. 2020; 394: 112112.
[89]
Wang S, Martin SM, Harris PS, Knudson CM. Caspase inhibition blocks cell death and enhances mitophagy but fails to promote T-cell lymphoma. PLoS ONE. 2011; 6: e19786.
[90]
Yamashima T. Hsp70.1 and related lysosomal factors for necrotic neuronal death. Journal of Neurochemistry. 2012; 120: 477–494.
[91]
Kim EH, Wong S, Martinez J. Programmed Necrosis and Disease:we interrupt your regular programming to bring you necroinflammation. Cell Death & Differentiation. 2019; 26: 25–40.
[92]
Liu Y, Liu T, Lei T, Zhang D, Du S, Girani L, et al. RIP1/RIP3-regulated necroptosis as a target for multifaceted disease therapy (Review). International Journal of Molecular Medicine. 2019; 44: 771–786.
[93]
Newton K. RIPK1 and RIPK3: critical regulators of inflammation and cell death. Trends in Cell Biology. 2015; 25: 347–353.
[94]
Kaczmarek A, Vandenabeele P, Krysko DV. Necroptosis: the release of damage-associated molecular patterns and its physiological relevance. Immunity. 2013; 38: 209–223.
[95]
Dunai Z, Bauer PI, Mihalik R. Necroptosis: biochemical, physiological and pathological aspects. Pathology Oncology Research. 2011; 17: 791–800.
[96]
Wallace HM, Pye KR. Necrosis. In: Schwab M, (ed) Encyclopedia of Cancer (pp. 1–5). Berlin, Heidelberg: Springer Berlin Heidelberg. 2016.
[97]
Bock FJ, Tait SWG. Mitochondria as multifaceted regulators of cell death. Nature Reviews Molecular Cell Biology. 2020; 21: 85–100.
[98]
Sesso A, Belizário JE, Marques MM, Higuchi ML, Schumacher RI, Colquhoun A, et al. Mitochondrial swelling and incipient outer membrane rupture in preapoptotic and apoptotic cells. The Anatomical Record. 2012; 295: 1647–1659.
[99]
Fulda S. Regulation of necroptosis signaling and cell death by reactive oxygen species. Biological Chemistry. 2016; 397: 657–660.
[100]
Hailey DW, Rambold AS, Satpute-Krishnan P, Mitra K, Sougrat R, Kim PK, et al. Mitochondria supply membranes for autophagosome biogenesis during starvation. Cell. 2010; 141: 656–667.
[101]
Lyamzaev KG, Tokarchuk AV, Panteleeva AA, Mulkidjanian AY, Skulachev VP, Chernyak BV. Induction of autophagy by depolarization of mitochondria. Autophagy. 2018; 14: 921–924.
[102]
Riaz TA, Junjappa RP, Handigund M, Ferdous J, Kim HR, Chae HJ. Role of endoplasmic reticulum stress sensor IRE1alpha in cellular physiology, calcium, ROS signaling, and metaflammation. Cells. 2020; 9: 1160.
[103]
Zhang Q, Liu J, Chen S, Liu J, Liu L, Liu G, et al. Caspase-12 is involved in stretch-induced apoptosis mediated endoplasmic reticulum stress. Apoptosis. 2016; 21: 432–442.
[104]
Ma X, Yu M, Hao C, Yang W. Shikonin induces tumor apoptosis in glioma cells via endoplasmic reticulum stress, and Bax/Bak mediated mitochondrial outer membrane permeability. Journal of Ethnopharmacology. 2020; 263: 113059.
[105]
Iurlaro R, Muñoz-Pinedo C. Cell death induced by endoplasmic reticulum stress. The FEBS Journal. 2016; 283: 2640–2652.
[106]
Yu L, Chen Y, Tooze SA. Autophagy pathway: Cellular and molecular mechanisms. Autophagy. 2018; 14: 207–215.
[107]
Appelqvist H, Waster P, Eriksson I, Rosdahl I, Ollinger K. Lysosomal exocytosis and caspase-8-mediated apoptosis in UVA-irradiated keratinocytes. Journal of Cell Science. 2013; 126: 5578–5584.
[108]
Alu A, Han X, Ma X, Wu M, Wei Y, Wei X. The role of lysosome in regulated necrosis. Acta Pharmaceutica Sinica B. 2020; 10: 1880–1903.
[109]
Sola-Riera C, Garcia M, Ljunggren HG, Klingstrom J. Hantavirus inhibits apoptosis by preventing mitochondrial membrane potential loss through up-regulation of the pro-survival factor BCL-2. PLoS Pathogens. 2020; 16: e1008297.
[110]
González‐Arzola K, Velázquez‐Cruz A, Guerra‐Castellano A, Casado‐Combreras MA, Pérez‐Mejías G, Díaz‐Quintana A, et al. New moonlighting functions of mitochondrial cytochromecin the cytoplasm and nucleus. FEBS Letters. 2019; 593: 3101–3119.
[111]
Mohamed MS, Bishr MK, Almutairi FM, Ali AG. Inhibitors of apoptosis: clinical implications in cancer. Apoptosis. 2017; 22: 1487–1509.
[112]
Mendez DL, Akey IV, Akey CW, Kranz RG. Oxidized or reduced cytochrome c and axial ligand variants all form the apoptosome in vitro. Biochemistry. 2017; 56: 2766–2769.
[113]
Bano D, Prehn JHM. Apoptosis-inducing factor (AIF) in physiology and disease: the tale of a repented natural born killer. EBioMedicine. 2018; 30: 29–37.
[114]
Kadam A, Mehta D, Jubin T, Mansuri MS, Begum R. Apoptosis inducing factor: cellular protective function in dictyostelium discoideum. Biochimica Et Biophysica Acta. 2020; 1861: 148158.
[115]
Zhang J, Ma G, Guo Z, Yu Q, Han L, Han M, et al. Study on the apoptosis mediated by apoptosis-inducing-factor and influencing factors of bovine muscle during postmortem aging. Food Chemistry. 2018; 266: 359–367.
[116]
Wiraswati HL, Hangen E, Sanz AB, Lam N, Reinhardt C, Sauvat A, et al. Apoptosis inducing factor (AIF) mediates lethal redox stress induced by menadione. Oncotarget. 2016; 7: 76496–76507.
[117]
Kratimenos P, Koutroulis I, Agarwal B, Theocharis S, Delivoria-Papadopoulos M. Effect of Src kinase inhibition on cytochrome c, Smac/DIABLO and apoptosis inducing factor (AIF) following cerebral hypoxia-ischemia in newborn piglets. Scientific Reports. 2017; 7: 16664.
[118]
Su J, Zhou L, Xia M, Xu Y, Xiang X, Sun L. Bcl-2 Family proteins are involved in the signal crosstalk between endoplasmic reticulum stress and mitochondrial dysfunction in tumor chemotherapy resistance. BioMed Research International. 2014; 2014: 234370.
[119]
Kavčič N, Pegan K, Turk B. Lysosomes in programmed cell death pathways: from initiators to amplifiers. Biological Chemistry. 2017; 398: 289–301.
[120]
Tardy C, Sabourdy F, Garcia V, Jalanko A, Therville N, Levade T, et al. Palmitoyl protein thioesterase 1 modulates tumor necrosis factor alpha-induced apoptosis. Biochimica et Biophysica Acta. 2009; 1793: 1250-1258.
[121]
Pišlar A, Perišić Nanut M, Kos J. Lysosomal cysteine peptidases - Molecules signaling tumor cell death and survival. Seminars in Cancer Biology. 2015; 35: 168–179.
[122]
Repnik U, Hafner Česen M, Turk B. Lysosomal membrane permeabilization in cell death: Concepts and challenges. Mitochondrion. 2014; 19: 49–57.
[123]
Llorens F, Thüne K, Sikorska B, Schmitz M, Tahir W, Fernández-Borges N, et al. Altered Ca2+ homeostasis induces calpain-cathepsin axis activation in sporadic creutzfeldt-jakob disease. Acta Neuropathologica Communications. 2017; 5: 35.
[124]
Smith MA, Schnellmann RG. Calpains, mitochondria, and apoptosis. Cardiovascular Research. 2012; 96: 32–37.
[125]
Guo C, Witter L, Rudolph S, Elliott H, Ennis K, Regehr W. Purkinje cells directly inhibit granule cells in specialized regions of the cerebellar cortex. Neuron. 2016; 91: 1330–1341.
[126]
Shimobayashi E, Kapfhammer JP. Increased biological activity of protein kinase C gamma is not required in Spinocerebellar ataxia 14. Molecular Brain. 2017; 10: 34.
[127]
Sarna JR, Hawkes R. Patterned Purkinje cell loss in the ataxic sticky mouse. The European Journal of Neuroscience. 2011; 34: 79–86.
[128]
Luo J. Mechanisms of ethanol-induced death of cerebellar granule cells. Cerebellum. 2012; 11: 145–154.
[129]
Benitez SG, Castro AE, Patterson SI, Muñoz EM, Seltzer AM. Hypoxic preconditioning differentially affects GABAergic and glutamatergic neuronal cells in the injured cerebellum of the neonatal rat. PLoS ONE. 2014; 9: e102056.
[130]
Au AK, Chen Y, Du L, Smith CM, Manole MD, Baltagi SA, et al. Ischemia-induced autophagy contributes to neurodegeneration in cerebellar Purkinje cells in the developing rat brain and in primary cortical neurons in vitro. Biochimica Et Biophysica Acta. 2015; 1852: 1902–1911.
[131]
Kandeel S, Elhosary NM, El-Noor MMA, Balaha M. Electric injury-induced purkinje cell apoptosis in rat cerebellum: histological and immunohistochemical study. Journal of Chemical Neuroanatomy. 2017; 81: 87–96.
[132]
Bell JD, Ai J, Chen Y, Baker AJ. Mild in vitro trauma induces rapid Glur2 endocytosis, robustly augments calcium permeability and enhances susceptibility to secondary excitotoxic insult in cultured Purkinje cells. Brain. 2007; 130: 2528–2542.
[133]
Roome CJ, Kuhn B. Dendritic coincidence detection in purkinje neurons of awake mice. Elife. 2020; 9: e59619.
[134]
Wilson AM, Schalek R, Suissa-Peleg A, Jones TR, Knowles-Barley S, Pfister H, et al. Developmental rewiring between cerebellar climbing fibers and purkinje cells begins with positive feedback synapse addition. Cell Reports. 2019; 29: 2849–2861.e6.
[135]
Balmer TS, Trussell LO. Selective targeting of unipolar brush cell subtypes by cerebellar mossy fibers. Elife. 2019; 8: e44964.
[136]
Chaumont J, Guyon N, Valera AM, Dugué GP, Popa D, Marcaggi P, et al. Clusters of cerebellar purkinje cells control their afferent climbing fiber discharge. Proceedings of the National Academy of Sciences of the United States of America. 2013; 110: 16223–16228.
[137]
Ryu C, Jang DC, Jung D, Kim YG, Shim HG, Ryu H, et al. STIM1 regulates somatic Ca2+ signals and intrinsic firing properties of cerebellar purkinje neurons. The Journal of Neuroscience. 2017; 37: 8876–8894.
[138]
Berridge MJ, Bootman MD, Roderick HL. Calcium signalling: dynamics, homeostasis and remodelling. Nature Reviews. Molecular Cell Biology. 2003; 4: 517–529.
[139]
Titley HK, Watkins GV, Lin C, Weiss C, McCarthy M, Disterhoft JF, et al. Intrinsic excitability increase in cerebellar purkinje cells after delay eye-blink conditioning in mice. The Journal of Neuroscience. 2020; 40: 2038–2046.
[140]
Kitamura K, Kano M. Dendritic calcium signaling in cerebellar purkinje cell. Neural Networks. 2013; 47: 11–17.
[141]
Empson RM, Turner PR, Nagaraja RY, Beesley PW, Knöpfel T. Reduced expression of the Ca2+ transporter protein PMCA2 slows Ca2+ dynamics in mouse cerebellar purkinje neurones and alters the precision of motor coordination. The Journal of Physiology. 2010; 588: 907–922.
[142]
Strehler EE, Thayer SA. Evidence for a role of plasma membrane calcium pumps in neurodegenerative disease: Recent developments. Neuroscience Letters. 2018; 663: 39–47.
[143]
De Miguel E, Álvarez-Otero R. Development of the cerebellum in turbot (psetta maxima): analysis of cell proliferation and distribution of calcium binding proteins. Journal of Chemical Neuroanatomy. 2017; 85: 60–68.
[144]
Attaai AH, Noreldin AE, Abdel-maksoud FM, Hussein MT. An updated investigation on the dromedary camel cerebellum (camelus dromedarius) with special insight into the distribution of calcium-binding proteins. Scientific Reports. 2020; 10: 21157.
[145]
Fierro L, Llano I. High endogenous calcium buffering in Purkinje cells from rat cerebellar slices. The Journal of Physiology. 1996; 496: 617–625.
[146]
Welsh JP, Yuen G, Placantonakis DG, Vu TQ, Haiss F, O’Hearn E, et al. Why do purkinje cells die so easily after global brain ischemia? Aldolase C, EAAT4, and the cerebellar contribution to posthypoxic myoclonus. Advances in Neurology. 2002; 89: 331–359.
[147]
Bäurle J, Kranda K, Frischmuth S. On the variety of cell death pathways in the lurcher mutant mouse. Acta Neuropathologica. 2006; 112: 691–702.
[148]
Selimi F, Doughty M, Delhaye-Bouchaud N, Mariani J. Target-related and intrinsic neuronal death in lurcher mutant mice are both mediated by caspase-3 activation. The Journal of Neuroscience. 2000; 20: 992–1000.
[149]
Wüllner U, Weller M, Schulz JB, Krajewski S, Reed JC, Klockgether T. Bcl-2, Bax and Bcl-x expression in neuronal apoptosis: a study of mutant weaver and lurcher mice. Acta Neuropathologica. 1998; 96: 233–238.
[150]
Yue Z, Horton A, Bravin M, DeJager PL, Selimi F, Heintz N. A novel protein complex linking the delta 2 glutamate receptor and autophagy: implications for neurodegeneration in lurcher mice. Neuron. 2002; 35: 921–933.
[151]
Erekat NS. Autophagy precedes apoptosis among at risk cerebellar purkinje cells in the shaker mutant rat: an ultrastructural study. Ultrastructural Pathology. 2018; 42: 162–169.
[152]
Erekat NS. Cerebellar Purkinje cells die by apoptosis in the shaker mutant rat. Brain Research. 2017; 1657: 323–332.
[153]
Cendelin J. From mice to men: lessons from mutant ataxic mice. Cerebellum & Ataxias. 2014; 1: 4.
[154]
Zanjani HS, McFarland R, Cavelier P, Blokhin A, Gautheron V, Levenes C, et al. Death and survival of heterozygous Lurcher Purkinje cells in vitro. Developmental Neurobiology. 2009; 69: 505–517.
[155]
Vernet-der Garabedian B, Derer P, Bailly Y, Mariani J. Innate immunity in the Grid2Lc/+ mouse model of cerebellar neurodegeneration: glial CD95/CD95L plays a non-apoptotic role in persistent neuron loss-associated inflammatory reactions in the cerebellum. Journal of Neuroinflammation. 2013; 10: 65.
[156]
Zanjani HS, Lohof AM, McFarland R, Vogel MW, Mariani J. Enhanced survival of wild-type and lurcher purkinje cells in vitro following inhibition of conventional PKCs or stress-activated MAP kinase pathways. Cerebellum. 2013; 12: 377–389.
[157]
Dumesnil-Bousez N, Sotelo C. Early development of the lurcher cerebellum: purkinje cell alterations and impairment of synaptogenesis. Journal of Neurocytology. 1992; 21: 506–529.
[158]
Vander Heiden MG, Chandel NS, Williamson EK, Schumacker PT, Thompson CB. Bcl-xL regulates the membrane potential and volume homeostasis of mitochondria. Cell. 1997; 91: 627–637.
[159]
Norman DJ, Feng L, Cheng SS, Gubbay J, Chan E, Heintz N. The lurcher gene induces apoptotic death in cerebellar Purkinje cells. Development. 1995; 121: 1183–1193.
[160]
Liang XH, Jackson S, Seaman M, Brown K, Kempkes B, Hibshoosh H, et al. Induction of autophagy and inhibition of tumorigenesis by beclin 1. Nature. 1999; 402: 672–676.
[161]
Lu W, Tsirka SE. Partial rescue of neural apoptosis in the Lurcher mutant mouse through elimination of tissue plasminogen activator. Development. 2002; 129: 2043–2050.
[162]
Armstrong CL, Duffin CA, McFarland R, Vogel MW. Mechanisms of compartmental purkinje cell death and survival in the lurcher mutant mouse. The Cerebellum. 2011; 10: 504–514.
[163]
Tolbert DL, Ewald M, Gutting J, La Regina MC. Spatial and temporal pattern of purkinje cell degeneration in shaker mutant rats with hereditary cerebellar ataxia. The Journal of Comparative Neurology. 1995; 355: 490–507.
[164]
Bertini E, des Portes V, Zanni G, Santorelli F, Dionisi-Vici C, Vicari S, et al. X-linked congenital ataxia: a clinical and genetic study. American Journal of Medical Genetics. 2000; 92: 53–56.
[165]
Figueroa KP, Paul S, Calì T, Lopreiato R, Karan S, Frizzarin M, et al. Spontaneous shaker rat mutant – a new model for X-linked tremor-ataxia. Disease Models & Mechanisms. 2016; 9: 553–562.
[166]
Tolbert DL, Clark BR. Olivocerebellar projections modify hereditary Purkinje cell degeneration. Neuroscience. 2000; 101: 417–433.
[167]
Tolbert DL, Clark BR. GDNF and IGF-i trophic factors delay hereditary Purkinje cell degeneration and the progression of gait ataxia. Experimental Neurology. 2003; 183: 205–219.
[168]
Marzella PL, Gillespie LN. Role of trophic factors in the development, survival and repair of primary auditory neurons. Clinical and Experimental Pharmacology & Physiology. 2002; 29: 363–371.
[169]
Desclin JC. Histological evidence supporting the inferior olive as the major source of cerebellar climbing fibers in the rat. Brain Research. 1974; 77: 365–384.
[170]
Hess BH, Krewet JA, Tolbert DL. Olivocerebellar projections are necessary for exogenous trophic factors to delay heredo-Purkinje cell degeneration. Brain Research. 2003; 986: 54–62.
[171]
Erekat NS. Active caspase-3 upregulation is augmented in at-risk cerebellar Purkinje cells following inferior olive chemoablation in the shaker mutant rat: an immunofluorescence study. Neurological Research. 2019; 41: 234–241.
[172]
Zanjani H, Rondi-Reig L, Vogel M, Martinou J, Delhaye-Bouchaud N, Mariani J. Overexpression of a hu-Bcl-2 transgene in lurcher mutant mice delays purkinje cell death. Comptes Rendus De L’AcadéMie Des Sciences - Series III - Sciences De La Vie. 1998; 321: 633–640.
[173]
Selimi F, Vogel MW, Mariani J. Bax inactivation in lurcher mutants rescues cerebellar granule cells but not purkinje cells or inferior olivary neurons. The Journal of Neuroscience. 2000; 20: 5339–5345.
[174]
Maycotte P, Guemez-Gamboa A, Moran J. Apoptosis and autophagy in rat cerebellar granule neuron death: Role of reactive oxygen species. Journal of Neuroscience Research. 2010; 88: 73–85.
[175]
Boya P, Gonzalez-Polo RA, Casares N, Perfettini JL, Dessen P, Larochette N, et al. Inhibition of macroautophagy triggers apoptosis. Molecular and Cellular Biology. 2005; 25: 1025–1040.
[176]
Piacentini M, Evangelisti C, Mastroberardino PG, Nardacci R, Kroemer G. Does prothymosin-alpha act as molecular switch between apoptosis and autophagy? Cell Death and Differentiation. 2003; 10: 937–939.
[177]
Wu YP, Mizukami H, Matsuda J, Saito Y, Proia RL, Suzuki K. Apoptosis accompanied by up-regulation of TNF-alpha death pathway genes in the brain of Niemann-Pick type C disease. Molecular Genetics and Metabolism. 2005; 84: 9–17.
Share
Back to top