Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3289 2022-07-21 10:03:40 |
2 update references and layout -6 word(s) 3283 2022-07-22 03:40:45 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Suárez-Rivero, J.M.;  Pastor-Maldonado, C.J.;  Povea-Cabello, S.;  Álvarez-Córdoba, M.;  Villalón-García, I.;  Talaverón-Rey, M.;  Suárez-Carrillo, A.;  Munuera-Cabeza, M.;  Reche-López, D.;  Cilleros-Holgado, P.; et al. Mitochondrial Unfolded Protein Response for Therapy. Encyclopedia. Available online: https://encyclopedia.pub/entry/25378 (accessed on 25 July 2024).
Suárez-Rivero JM,  Pastor-Maldonado CJ,  Povea-Cabello S,  Álvarez-Córdoba M,  Villalón-García I,  Talaverón-Rey M, et al. Mitochondrial Unfolded Protein Response for Therapy. Encyclopedia. Available at: https://encyclopedia.pub/entry/25378. Accessed July 25, 2024.
Suárez-Rivero, Juan M., Carmen J. Pastor-Maldonado, Suleva Povea-Cabello, Mónica Álvarez-Córdoba, Irene Villalón-García, Marta Talaverón-Rey, Alejandra Suárez-Carrillo, Manuel Munuera-Cabeza, Diana Reche-López, Paula Cilleros-Holgado, et al. "Mitochondrial Unfolded Protein Response for Therapy" Encyclopedia, https://encyclopedia.pub/entry/25378 (accessed July 25, 2024).
Suárez-Rivero, J.M.,  Pastor-Maldonado, C.J.,  Povea-Cabello, S.,  Álvarez-Córdoba, M.,  Villalón-García, I.,  Talaverón-Rey, M.,  Suárez-Carrillo, A.,  Munuera-Cabeza, M.,  Reche-López, D.,  Cilleros-Holgado, P.,  Piñero-Pérez, R., & Sánchez-Alcázar, J.A. (2022, July 21). Mitochondrial Unfolded Protein Response for Therapy. In Encyclopedia. https://encyclopedia.pub/entry/25378
Suárez-Rivero, Juan M., et al. "Mitochondrial Unfolded Protein Response for Therapy." Encyclopedia. Web. 21 July, 2022.
Mitochondrial Unfolded Protein Response for Therapy
Edit

The mitochondrial unfolded protein response (UPRmt) is a mechanism aimed to preserve or repair damaged mitochondria. UPRmt is responsible for maintaining mitochondrial proteostasis via mitochondrial activation of a transcriptional program in the nuclear DNA. This compensation system can be divided into three main pathways: activation of (1) chaperones, which boost refolding of misfolded proteins to restore them to their functional conformation and assist the folding of newly synthesized proteins; (2) proteases that are able to degrade aberrant proteins or aggregates; and (3) an antioxidant system that palliates ROS overproduction. Although human UPRmt is not fully understood, it is gaining relevance in a variety of physiological processes on top of its canonical function, such as ageing, oxidative stress resistance, hematopoietic stem cell maintenance, glycolysis, antibacterial immunity, coenzyme Q biosynthesis, and mitochondrial fission. Loss of mitochondrial proteostasis is the main UPRmt inducer. Accumulation of damaged proteins exceeding the protein-processing capacity of the chaperones and proteases in mitochondria would activate UPRmt, for instance. Additionally, factors interfering with mitochondrial function promote UPRmt induction. Examples of these are the inhibition of complex I by rotenone, bacterial toxins, knockdown of quality control proteins, or generation of excess ROS by paraquat.

mitochondria homeostasis unfolded protein response therapeutic target

1.UPRmt in Mitochondrial Diseases

Mitochondrial diseases encompass a broad spectrum of muscular and neurodegenerative disorders, both chronic and progressive, caused by mutations in nuclear (nDNA) or mitochondrial (mtDNA) DNA [1]. The prevalence of these diseases has been established at 1:5000 [2]; however, this number is increasing due to the standardization of whole-exome sequencing [3]. Most OXPHOS disorders in children are a consequence of the mutation of nuclear DNA, and they are transmitted as autosomal recessive traits, usually with severe phenotypes and a fatal outcome. Among the maternally inherited pathogenic mtDNA mutations, more than 50% have been identified in genes encoded by mitochondrial transfer RNAs (mt-tRNA) (MTT genes). At first sight, it might seem paradoxical to induce a mild mitochondria stress in a general mitochondrial dysfunction context; however, recent studies present this strategy as a new alternative treatment for mitochondrial diseases.
Perry et al. showed that the tetracycline antibiotics family increased cell survival and fitness in MELAS cybrids and Rieske cells (Knockout Complex III mouse fibroblasts) under glucose restriction [4]. Specifically, doxycycline improved survival in wild-type cells treated with piericidin (complex I inhibitor) or antimycin (complex III inhibitor) during glucose deprivation. In addition to tetracyclines, the anti-parasitic agent pentamidine and the antibiotic retapamulin also scored positive on the screening in MELAS and mutant ND1 cybrid cells. They propose a “mitohormetic effect” by the induction of ATF4 (UPRmt) and p-eIF2α (ISR) as mechanism of action; however, in some cases, cell survival was independent of ATF4, suggesting the high complexity of this pathway. Suarez-Rivero et al. also demonstrated that tetracycline treatment boosts the production of UPRmt-related proteins and promotes the activation of pathways involving cAMP and cGMP, which might be implicated in mitochondrial compensatory mechanisms comprising sirtuins and chaperones’ activity [5][6][7]. Tetracycline treatment and the subsequent activation of UPRmt, the increased number of chaperones, and mitochondrial auxiliary proteins may promote the stability of a fraction of mutated mitochondrial proteins which would carry out their function to some extent [8]. All of these antibiotics have one factor in common: they are inhibitors of the mitochondrial translation at different degrees [9][10][11].
Given the bacterial origin of mammalian cells’ mitochondria and the fact that they conserve prokaryotic features such as the 55S or 60S ribosomes, it is understandable that mitochondria are exceptionally sensitive to antibiotics [12]. In this way, mitohormesis can be triggered by a partial inhibition of mitochondrial translation promoting a beneficial retrograde signaling response including the modulation of mitochondrial dynamics, the expression of nuclear and mitochondrial-encoded genes, the antioxidant response, stimulating mitochondrial function, and boosting cellular defense mechanisms that increase stress resistance [13].
Although the research of Perry et al. and Suarez-Rivero et al. is promising, antibiotic use in mitochondrial diseases is still in debate [14]. Several antibiotics can worsen the conditions of individuals bearing mtDNA mutations. Aminoglycosides, for instance, induce ototoxic hearing loss in subjects with mutations in the 12S rRNA gene [15]. Although aminoglycosides typically display specificity toward prokaryotes over eukaryotes, human mitochondrial ribosomes that have A1408 and G1491 at analogous positions exhibit higher resemblance to their bacterial counterparts. This similarity is likely responsible for some of the adverse effects shown by aminoglycosides [16].
It has recently been demonstrated that pterostilbene in combination with mitochondrial cofactors treatment activates SIRT3 and UPRmt as compensatory mechanisms, as well as enhances sirtuins’ levels and mitochondrial activity in several cell models of mitochondrial diseases [17]. In contrast to long-term antibiotic treatment, pterostilbene is considered safe for human consumption and presents numerous well-known features, such as a prominent antioxidant activity and a high anti-inflammatory potential [18]. Furthermore, it has been reported to prolong lifespan in several animal models [19] due to its neuroprotective [20] and cardioprotective [21] properties. At the molecular level, pterostilbene activates sirtuins [19] and has been linked to AMP-activated protein kinase (AMPK) [22] and Nuclear factor erythroid 2-related factor 2 (Nrf2) by recent studies [23], suggesting that it could be a promising compound to maintain mitochondrial homeostasis.

2. UPRmt in Neurodegeneration

Neurodegenerative diseases are a widely heterogeneous group of disorders characterized by the progressive degeneration of the structure and function of the central and peripheral nervous system. Although they affect millions of patients worldwide and have a profound impact on families and their community, the pathogenic mechanisms underlying these conditions remain unclear. The therapeutic options for patients are scarce and merely palliative. Nevertheless, there is currently great interest among the scientific community in regard to the identification of biomarkers or concrete genetic mutations that might help clinicians anticipate the onset of these diseases in a way to either treat them when still reversible or slow down their development [24]. This is especially relevant given the fact that the protein abnormalities that are characteristic of these diseases are present in patients long before the clinical symptoms become noticeable [25][26].
Most neurodegenerative diseases share in common the accumulation of misfolded proteins; however, they also possess traits associated with progressive neuronal dysfunction and death, such as proteotoxic stress, abnormalities in autophagy, neuroinflammation, apoptosis, ROS production, and mitochondrial dysfunction [27]. Mitochondrial function is of utmost importance for neurons because limited glycolysis causes them to rely exclusively on OXPHOS for energy production. Given that the long neuronal axons require energy transport over long distances, and because synaptic transmission is dependent on calcium signaling, mitochondrial performance needs to be tightly regulated [28]. Taken together, there is compelling evidence to suggest a relevant mitochondrial involvement in the pathogenesis of several neurodegenerative diseases, with their role being particularly significant in Parkinson’s disease (PD). According to several studies, patients of this condition present a selective deficiency of respiratory chain complex I, which is most remarkable in the substantia nigra [29][30]. In Alzheimer’s diseases (AD), it has been demonstrated that patients present impaired oxygen consumption in the brain, further adding to the notion that bioenergetic dysfunction and mitochondrial impairment are common features [31]. In fact, impairment of several mitochondrial enzymes has been detected in AD patients, where respiratory chain complex IV is most affected [32]. Complex IV activity was found to be significantly reduced in the brain tissue of AD patients, which presented a general decrease of electron transport chain function [33]. Apart from AD and PD, mitochondrial perturbations have been reported in many other neurodegenerative diseases, among which amyotrophic lateral sclerosis (ALS) and Huntington’s disease (HD) deserve special mention [34]. Nevertheless, a critical question that remains unanswered is whether mitochondrial dysfunction contributes to the initial stages of neurodegeneration, being primarily responsible for the onset of pathogenesis or just a secondary feature arising from alternative phenomena such as the accumulation of misfolded proteins or cellular stress. Either way, it has been demonstrated that targeting mitochondrial dysfunction is a promising therapeutic strategy for neurodegenerative conditions [35].

2.1. Parkinson’s Diseases

PD is an age-associated neurodegenerative movement disorder that is mainly caused by the death of dopaminergic neurons in the brain substantia nigra [36]. UPRmt may regulate the occurrence and development of PD. It has been shown that a PINK1 mutation can activate ATFS-1-dependent UPRmt and promote dopaminergic neuron survival in a PD worm model [37]. Ginseng protein protected against neurodegeneration by inducing UPRmt in a PINK1 fly model of PD [38]. Dastidar SG et al. demonstrated that activation of 4E-BP1 correlates with UPRmt induction, which reduces PD associated neurotoxicity in mouse neurons [39]. In fact, mutant LRRK2G2019S, the most common PD-causing allele in humans, results in reduced 4E-BP1 function and may contribute to PD pathogenesis by UPRmt underactivation [40]. In addition, altered omi and HtrA2 protein, UPRmt-related proteases, causes neuropathy with the same characteristics as PD [41].

2.2. Alzheimer’s Disease

Alzheimer’s disease (AD) is the most common form of dementia [42]. Clinically, AD is defined by cognitive impairment that is pervasive enough to interfere with a person’s ability to work or complete daily activities. The main histologic characteristics include amyloid plaques that largely contain amyloid beta (Aβ) protein and neurofibrillary tangles (NFTs) that consist of hyperphosphorylated tau protein, both of which cause premature neuronal death. Current AD treatments confer a slight benefit but ultimately do not prevent progression [43].
Mitochondria in AD show numerous changes [44]. First, complex-IV-reduced activity has consistently been observed across numerous tissues. In brain tissue specifically, the decrease in complex IV subunits has been linked with disease progression [45]. Mitochondrial surface area also decreases, cristae structures are altered, and an increased variability in mitochondrial shape has been observed [46]. There are changes in fusion and fission mitochondrial proteins that appear to affect mitochondrial localization in AD neurons [47]. Indeed, neuronal cultures with AD-relevant fission/fusion protein alterations recapitulate AD-like changes in mitochondrial distribution [48].
Compared to cognitively intact controls, the expression of UPRmt-related genes is increased by 40–60% in sporadic AD subjects and 70–90% in familial AD, respectively [49]. Sorrentino et al. showed that increasing mitochondrial proteostasis by targeting mitochondrial translation and mitophagy both pharmacologically and genetically increases the fitness and lifespan of worms, cells, and mouse models of AD and reduces amyloid aggregation [50]. In addition, UPRmt is strongly activated and exerts a protective role against Aβ protein toxicity in PITRM1-knockout human cells. In line with this, pharmacological inhibition of UPRmt exacerbates the Aβ proteotoxicity in PITRM1-knockout human cells [51]. Herein, UPRmt is activated in the mouse brains and human SHSY5Y cells after Aβ treatment, while the inhibition of UPRmt aggravates cytotoxic effects of Aβ [52]. Treatments which increase the NAD+ pool, a sirtuin cofactor, alleviate protein toxicity and improve the memory of AD mouse by activating the UPRmt. Taken together, these studies provide the evidence of the protective role of UPRmt on AD. Furthermore, the activation of UPRmt induced by Aβ may depend on the regulation of two signaling pathways: the mevalonic acid pathway and ceramide pathway [53]. Therefore, the activation of UPRmt can relieve the symptoms of AD.

2.3. Huntington’s Disease

Huntington’s disease (HD) is a fatal and inherited neurodegenerative disorder that progresses for 15–20 years after the initial onset [54]. The main cause of HD is the expanded CAG repeats encoding polyglutamine (polyQ) in the N-terminus of the huntingtin (Htt) protein [55]. The genetic cause of HD was identified more than 20 years ago. However, the underlying mechanisms leading to the pathogenesis of HD remain elusive.
Accumulating evidence suggests that mitochondrial dysfunction plays an important role in the pathogenesis of HD [56]. For instance, mutant Htt associates with the outer mitochondrial membrane in different HD models, resulting in mitochondrial permeability transition pore opening, calcium disturbance, reduced ATP production, mitochondrial membrane potential loss, increased ROS production, and premature release of cytochrome c [57]. How mutant Htt can affect selectively medium spiny striatal neurons despite its ubiquitous expression is a key question that still remains unanswered. One of the several theories that have been proposed points again to mitochondria, suggesting that medium spiny neurons, characterized by particularly high-energy demands, are especially susceptible to mutant Huntingtin-induced mitochondrial dysfunction and inhibition of respiration [58].
Fu et al. [59] found that mutant Htt suppressed the expression of ABCB10, a component of the UPRmt [60], in various HD models by impairing its mRNA stability. Deletion of ABCB10 induced ROS production and cell death in HD mouse striatal cells. Moreover, ABCB10 was required for CHOP activation under mitochondrial stress. They also showed that chaperone HSP60 and protease Clpp, two downstream genes of CHOP [61], were decreased in HD cells.

2.4. Amyotrophic Lateral Sclerosis

Amyotrophic lateral sclerosis (ALS) is caused by the progressive degeneration of motor neurons in the spinal cord, the brain stem, and the motor cortex. Motor neuron death prompts muscle weakness and paralysis, causing death in 1–5 years from the time of symptoms’ onset. The primary identified ALS-linked gene was superoxide dismutase 1 (SOD1), an antioxidant protein [62]. However, mutations in several other genes have been reported, and various molecular pathways have been associated to neuronal death in ALS [63]. The only two disease-modifying potential therapies currently approved for ALS are riluzole and edaravone, but the exact neuroprotective action of these drugs is unknown, and they present no obvious improvement in ALS patients’ health [64]. The absence of alternative drugs for the treatment of ALS indicates the need for the implementation of novel therapeutic strategies
ALS-associated mitochondrial dysfunction comes in many shapes, including defective OXPHOS, ROS production, impaired calcium buffering capacity, and defective mitochondrial dynamics. In addition, mitochondrial dysfunction appears to be directly or indirectly linked to all mechanisms of toxicity associated with ALS, including excitotoxicity, loss of protein homeostasis, and defective axonal transport [65].
TDP-43 proteinopathy is characterized by the presence of TDP-43 immunoreactive inclusion bodies in the affected tissues. This alteration is present in several neurodegenerative diseases with high predominance in ALS [66][67]. Peng et al. showed that LonP1, one of the key mitochondrial proteases constituting the UPRmt response, protects against TDP-43-induced cytotoxicity and neurodegeneration. They suggest the activation of LonP1 via UPRmt as a treatment for TDP-43 proteinopathy, protecting against or reversing mitochondrial damage as a potential therapeutic approach to these neurodegenerative disorders [68]. Qi et al. demonstrated that the mitochondrial function pathway was disrupted in the brain of SOD1G93A mice, and the replenishment of intracellular NAD+, by providing nicotinamide, could reduce neurotoxic protein aggregates of mitochondria in the brain of SOD1G93A mice [69]. Moreover, they concluded that nicotinamide might modulate mitochondrial proteostasis and improve adult neurogenesis through activation of the UPRmt signaling in the brain of SOD1G93A mice. Curiously, it has been reported that UPRmt is transiently activated in the spinal cord of ALS mice models in the late symptomatic phase. However, there is a significant sex difference in the activation of the UPRmt: it was significantly activated in female SOD1G93A mice but not in males [70]. Finally, Gomez et al. hypothesized that SOD1 may have multiple functions: antioxidant, UPRmt regulator, and gene transcription enhancer [71]. Taken together, these functions place SOD1 as a key regulator of the communication between the nucleus and mitochondria.

3. UPRmt in Heart Diseases

Maintaining mitochondrial quality in cardiomyocytes is essential, given that they produce 8% of the total ATP consumed by the organism and power the constant contraction and relaxation of the myocardium [72]. In fact, mitochondrial abnormalities are a common feature to all types of cardiomyopathies [73]. Mitochondria with aberrant structures are commonly observed in cardiac cells of all forms of heart disease [74]. Furthermore, cardiomyopathy has been reported due to mitochondrial disease [75], and mitochondrial dysfunction has been linked with hypertension. It is nowadays clear that the importance of functional mitochondria for cardiac health is undeniable. One of the main mechanisms ensuring mitochondrial fitness is proteostasis. It finetunes biogenesis, folding, and degradation of mitochondrial proteins, processes which are commonly altered during cardiac stress. It has been demonstrated that stimulation of UPRmt improves mitochondrial function and reduces cardiac damage in response to stress. In this regard, a recent study proved that UPRmt enhancement with small-molecule agents ameliorates mitochondrial and contractile dysfunction in the murine heart. This is the reason why UPRmt  was proposed as a potential therapeutic target in heart failure [76].
Wang et al. demonstrated the relevance of mitophagy and UPRmt in myocardial injuries and stress [77]. They proposed UPRmt activation with oligomycin, a complex V inhibitor, as a mechanism to reduce sepsis-mediated mitochondrial injury and myocardial dysfunction; however, this cardioprotective effect was imperceptible in mitophagy-inhibited mice. On the other hand, when UPRmt was inhibited, mitophagy-mediated protection of mitochondria and cardiomyocytes was partly blunted. Taken together, their observations suggest that both UPRmt and mitophagy are slightly activated by myocardial stress and that they work together to sustain mitochondrial performance and cardiac function.

4. UPRmt in Lifespan

Upregulation of UPRmt has been proposed as the common pathway in lifespan extension induced by mitochondrial defects [78]. Studies in C. elegans revealed that the induction of mitochondrial stress at an early age resulted in lifespan extension and UPRmt upregulation which persisted even after the stress inductor was withdrawn [79]. This observation suggests that UPRmt upregulation is directly responsible for lifespan extension in worms. Supporting this idea, Yokoyama et al. demonstrated that transgenic worms expressing the nematode UPRmt protein Hsp70F live over 40% more than wild-type animals [80]. To prove whether UPRmt upregulation also influenced lifespan extension in mammals, a recent study assessed UPRmt activation in a long-lived mouse model, Snell dwarf mice. These mice have a single point mutation in Pit1 and show more than 40% increase in mean and maximal lifespan [81]. The results show that UPRmt was upregulated in both living mice and cultured cells. Moreover, it was reported that elevated basal and post-stress UPRmt levels in primary fibroblasts derived from Snell mice. However, at the organismal level, UPRmt was upregulated only after doxycycline-induced mitochondrial stress exposure [82].
The ability of UPRmt activation to prolong lifespan can be explained by several reasons: its positive impact on mitochondrial function, its ability to reduce mitochondrial metabolic by-products, its modulation of insulin/IGF and mTOR signaling, and its involvement in hormetic cellular maintenance [83]. Following UPRmt activation, mitochondria undergo structural and functional changes. The mitochondrial network becomes fragmented [78], and cellular oxygen consumption is significantly reduced [84][85]. This particular phenotype may be a coordinated response to stress that favors mitophagy and lowers respiration to enable mitochondrial repair. Taken together, these measures would ultimately lead to enhanced homeostasis and cell survival, consequently increasing lifespan.
On top of this, UPRmt is responsible for changes in the production of a series of metabolic by-products with relevant physiological functions. Amidst these, ROS is the most widely studied. It elicits varied cellular stress responses, and by doing so, it influences homeostasis and longevity [86]. Additionally, the UPRmt has an impact on the mevalonate pathway, an important target of statins, and on NAD+, which is the substrate for poly (ADP-ribose) polymerases (PARPs) and sirtuin deacetylases and acts as major longevity regulator [87][88]. The link of UPRmt to longevity might also be explained by its effect on insulin signaling. Reduced insulin/IGF-1 signaling has been directly linked to a series of protective mechanisms such as regulation of endoplasmic reticulum unfolded protein response (UPRER) or activation of heat shock factor 1 (HIF-1), which boosts the expression of small heat shock proteins, leading to lifespan extension in worms [89]. In this line, it has been proven that impairment of the insulin/IGF-1 signaling induces a transient ROS stress signal that potentially activates UPRmt and results in prolonged lifespan in C. elegans [90]. Improving the level of cellular NAD+ in mice not only increases mitochondrial function but also induces the expression of UPRmt-associated genes, preventing skeletal muscle stem cells from aging and enhancing life span [91].
Although several different routes have been proposed to explain UPRmt implication in longevity, there is the possibility that its lifespan-prolonging ability is the result of improved cellular housekeeping. It is extensively known that loss of proteostasis and the consequent accumulation of misfolded proteins are a fundamental cause of ageing. UPRmt activation would palliate cellular damage in this context by restoring protein quality control. Not only does it assist correct protein folding via enhancement of chaperones’ expression, but it also promotes refolding of aberrant products and sequestration of protein aggregates in less cytotoxic states [92].

References

  1. Zeviani, M.; Carelli, V. Mitochondrial disorders. Curr. Opin. Neurol. 2007, 20, 564–571.
  2. Gorman, G.S.; Schaefer, A.M.; Ng, Y.; Gomez, N.; Blakely, E.L.; Alston, C.L.; Feeney, C.; Horvath, R.; Yu-Wai-Man, P.; Chinnery, P.F.; et al. Prevalence of nuclear and mitochondrial DNA mutations related to adult mitochondrial disease. Ann. Neurol. 2015, 77, 753–759.
  3. Molnar, M.J.; Kovacs, G.G. Mitochondrial diseases. Handb. Clin. Neurol. 2017, 145, 147–155.
  4. Perry, E.A.; Bennett, C.F.; Luo, C.; Balsa, E.; Jedrychowski, M.; O’Malley, K.E.; Latorre-Muro, P.; Ladley, R.P.; Reda, K.; Wright, P.M.; et al. Tetracyclines promote survival and fitness in mitochondrial disease models. Nat. Metab. 2021, 3, 33–42.
  5. Zhang, F.; Zhang, L.; Qi, Y.; Xu, H. Mitochondrial cAMP signaling. Cell. Mol. Life Sci. 2016, 73, 4577–4590.
  6. Scarpulla, R.C.; Vega, R.B.; Kelly, D.P. Transcriptional integration of mitochondrial biogenesis. Trends Endocrinol. Metab. 2012, 23, 459–466.
  7. Ventura-Clapier, R.; Garnier, A.; Veksler, V. Transcriptional control of mitochondrial biogenesis: The central role of PGC-1 alpha. Cardiovasc. Res. 2008, 79, 208–217.
  8. Suárez-Rivero, J.M.; Pastor-Maldonado, C.J.; Povea-Cabello, S.; Álvarez-Córdoba, M.; Villalón-García, I.; Talaverón-Rey, M.; Suárez-Carrillo, A.; Munuera-Cabeza, M.; Reche-López, D.; Cilleros-Holgado, P.; et al. UPR(mt) activation improves pathological alterations in cellular models of mitochondrial diseases. Orphanet J. Rare Dis. 2022, 17, 204.
  9. Chukwudi, C.U. rRNA binding sites and the molecular mechanism of action of the tetracyclines. Antimicrob. Agents Chemother. 2016, 60, 4433–4441.
  10. Meydan, S.; Marks, J.; Klepacki, D.; Sharma, V.; Baranov, P.V.; Firth, A.E.; Margus, T.; Kefi, A.; Vázquez-Laslop, N.; Mankin, A.S. Retapamulin-assisted ribosome profiling reveals the alternative bacterial proteome. Mol. Cell 2019, 74, 481–493.
  11. Sun, T.; Zhang, Y. Pentamidine binds to tRNA through non-specific hydrophobic interactions and inhibits aminoacylation and translation. Nucleic Acids Res. 2008, 36, 1654–1664.
  12. Wang, X.; Ryu, D.; Houtkooper, R.H.; Auwerx, J. Antibiotic use and abuse: A threat to mitochondria and chloroplasts with impact on research, health, and environment. BioEssays 2015, 37, 1045–1053.
  13. Palmeira, C.; Teodoro, J.; Amorim, J.A.; Steegborn, C.; Sinclair, D.; Rolo, A.P. Mitohormesis and metabolic health: The interplay between ROS, cAMP and sirtuins. Free Radic. Biol. Med. 2019, 141, 483–491.
  14. Suárez-Rivero, J.M.; Pastor-Maldonado, C.J.; Povea-Cabello, S.; Álvarez-Córdoba, M.; Villalón-García, I.; Talaverón-Rey, M.; Suárez-Carrillo, A.; Munuera-Cabeza, M.; Sánchez-Alcázar, J.A. Mitochondria and antibiotics: For good or for evil? Biomolecules 2021, 11, 1050.
  15. Gao, Z.; Chen, Y.; Guan, M.-X. Mitochondrial DNA mutations associated with aminoglycoside induced ototoxicity. J. Otol. 2017, 12, 1–8.
  16. McCoy, L.S.; Xie, Y.; Tor, Y. Antibiotics that target protein synthesis. Wiley Interdiscip. Rev. RNA 2011, 2, 209–232.
  17. Suárez-Rivero, J.M.; Pastor-Maldonado, C.J.; Romero-González, A.; Gómez-Fernandez, D.; Povea-Cabello, S.; Álvarez-Córdoba, M.; Villalón-García, I.; Talaverón-Rey, M.; Suárez-Carrillo, A.; Munuera-Cabeza, M.; et al. Pterostilbene in combination with mitochondrial cofactors improve mitochondrial function in cellular models of mitochondrial diseases. Front. Pharmacol. 2022, 13, 862085.
  18. Lange, K.W.; Li, S. Resveratrol, pterostilbene, and dementia. Biofactors 2018, 44, 83–90.
  19. Li, Y.-R.; Li, S.; Lin, C.-C. Effect of resveratrol and pterostilbene on aging and longevity. BioFactors 2018, 44, 69–82.
  20. Li, Q.; Chen, L.; Liu, X.; Li, X.; Cao, Y.; Bai, Y.; Qi, F. Pterostilbene inhibits amyloid-beta-induced neuroinflammation in a microglia cell line by inactivating the NLRP3/caspase-1 inflammasome pathway. J. Cell Biochem. 2018, 119, 7053–7062.
  21. Natalin, H.M.; Garcia, A.F.E.; Ramalho, L.N.Z.; Restini, C.B.A. Resveratrol improves vasoprotective effects of captopril on aortic remodeling and fibrosis triggered by renovascular hypertension. Cardiovasc. Pathol. 2016, 25, 116–119.
  22. Malik, S.A.; Acharya, J.D.; Mehendale, N.K.; Kamat, S.S.; Ghaskadbi, S.S. Pterostilbene reverses palmitic acid mediated insulin resistance in HepG2 cells by reducing oxidative stress and triglyceride accumulation. Free Radic. Res. 2019, 53, 815–827.
  23. Zhou, J.; Ci, X.; Ma, X.; Yu, Q.; Cui, Y.; Zhen, Y.; Li, S. Pterostilbene Activates the Nrf2-Dependent Antioxidant Response to Ameliorate Arsenic-Induced Intracellular Damage and Apoptosis in Human Keratinocytes. Front. Pharmacol. 2019, 10, 497.
  24. Beach, T.G. A review of biomarkers for neurodegenerative disease: Will they swing us across the valley? Neurol. Ther. 2017, 6 (Suppl. 1), 5–13.
  25. Evidente, V.G.H.; Adler, C.H.; Sabbagh, M.N.; Connor, D.J.; Hentz, J.G.; Caviness, J.N.; Sue, L.I.; Beach, T.G. Neuropathological findings of PSP in the elderly without clinical PSP: Possible incidental PSP? Parkinsonism Relat. Disord. 2011, 17, 365–371.
  26. Frigerio, R.; Fujishiro, H.; Ahn, T.B.; Josephs, K.A.; Maraganore, D.M.; DelleDonne, A.; Parisi, J.E.; Klos, K.J.; Boeve, B.F.; Disckson, D.W.; et al. Incidental Lewy body disease: Do some cases represent a preclinical stage of dementia with Lewy bodies? Neurobiol. Aging 2011, 32, 857–863.
  27. Katsnelson, A.; De Strooper, B.; Zoghbi, H.Y. Neurodegeneration: From cellular concepts to clinical applications. Sci. Transl. Med. 2016, 8, 364.
  28. Burchell, V.S.; Gandhi, S.; Deas, E.; Wood, N.; Abramov, A.; Plun-Favreau, H. Targeting mitochondrial dysfunction in neurodegenerative disease: Part I. Expert Opin. Ther. Targets 2010, 14, 369–385.
  29. Schapira, A.H.V.; Cooper, J.M.; Dexter, D.; Clark, J.B.; Jenner, P.; Marsden, C.D. Mitochondrial complex I deficiency in Parkinson’s disease. Lancet 1989, 333, 1269.
  30. Schapira, A.H.V.; Mann, V.M.; Cooper, J.M.; Dexter, D.; Daniel, S.E.; Jenner, P.; Clark, J.B.; Marsden, C.D. Anatomic and Disease Specificity of NADH CoQ1 Reductase (Complex I) Deficiency in Parkinson’s Disease. J. Neurochem. 1990, 55, 2142–2145.
  31. Tohgi, H.; Yonezawa, H.; Takahashi, S.; Sato, N.; Kato, E.; Kudo, M.; Hatano, K.; Sasaki, T. Cerebral blood flow and oxygen metabolism in senile dementia of Alzheimer’s type and vascular dementia with deep white matter changes. Neuroradiology 1998, 40, 131.
  32. Musatov, A.; Robinson, N.C. Susceptibility of mitochondrial electron-transport complexes to oxidative damage. Focus on cytochrome c oxidase. Free Radic. Res. 2012, 46, 1313–1326.
  33. Parker, W.D.; Parks, J.; Filley, C.M.; Kleinschmidt-DeMasters, B.K. Electron transport chain defects in Alzheimer’s disease brain. Neurology 1994, 44, 1090.
  34. Lezi, E.; Swerdlow, R.H. Mitochondria in neurodegeneration. Adv. Exp. Med. Biol. 2012, 942, 269–286.
  35. Bader, V.; Winklhofer, K.F. Mitochondria at the interface between neurodegeneration and neuroinflammation. Semin. Cell Dev. Biol. 2020, 99, 163–171.
  36. Puspita, L.; Chung, S.Y.; Shim, J.-W. Oxidative stress and cellular pathologies in Parkinson’s disease. Mol. Brain 2017, 10, 53.
  37. Cooper, J.F.; Machiela, E.; Dues, D.J.; Spielbauer, K.K.; Senchuk, M.M.; Van Raamsdonk, J.M. Activation of the mitochondrial unfolded protein response promotes longevity and dopamine neuron survival in Parkinson’s disease models. Sci. Rep. 2017, 7, 16441.
  38. Liu, M.; Yu, S.; Wang, J.; Qiao, J.; Liu, Y.; Wang, S.; Zhao, Y. Ginseng protein protects against mitochondrial dysfunction and neurodegeneration by inducing mitochondrial unfolded protein response in Drosophila melanogaster PINK1 model of Parkinson’s disease. J. Ethnopharmacol. 2020, 247, 112213.
  39. Dastidar, S.G.; Pham, M.T.; Mitchell, M.B.; Yeom, S.G.; Jordan, S.; Chang, A.; Sopher, B.L.; La Spada, A.R. 4E-BP1 Protects neurons from misfolded protein stress and Parkinson’s disease toxicity by inducing the mitochondrial unfolded protein response. J. Neurosci. 2020, 40, 8734–8745.
  40. Imai, Y.; Gehrke, S.; Wang, H.-Q.; Takahashi, R.; Hasegawa, K.; Oota, E.; Lu, B. Phosphorylation of 4E-BP by LRRK2 affects the maintenance of dopaminergic neurons in Drosophila. EMBO J. 2008, 27, 2432–2443.
  41. Strauss, K.M.; Martins, L.M.; Plun-Favreau, H.; Marx, F.P.; Kautzmann, S.; Berg, D.; Gasser, T.; Wszolek, Z.; Müller, T.; Bornemann, A.; et al. Loss of function mutations in the gene encoding Omi/HtrA2 in Parkinson’s disease. Hum. Mol. Genet. 2005, 14, 2099–3111.
  42. 2020 Alzheimer’s disease facts and figures. Alzheimers Dement. 2020, 16, 391–460.
  43. Grossberg, G.T.; Tong, G.; Burke, A.D.; Tariot, P.N. Present algorithms and future treatments for Alzheimer’s disease. J. Alzheimer’s Dis. 2019, 67, 1157–1171.
  44. Swerdlow, R.H. Mitochondria and cell bioenergetics: Increasingly recognized components and a possible etiologic cause of Alzheimer’s disease. Antioxid. Redox Signal 2012, 16, 1434–1455.
  45. Kish, S.J.; Mastrogiacomo, F.; Guttman, M.; Furukawa, Y.; Taanman, J.W.; Dozic, S.; Pandolfo, M.; Lamarche, J.; DiStefano, L.; Chang, L.-J. Decreased brain protein levels of cytochrome oxidase subunits in Alzheimer’s disease and in hereditary spinocerebellar ataxia disorders: A nonspecific change? J. Neurochem. 1999, 72, 700–707.
  46. Baloyannis, S.J. Mitochondrial alterations in Alzheimer’s disease. J. Alzheimers Dis. 2006, 9, 119–126.
  47. Manczak, M.; Calkins, M.J.; Reddy, P.H. Impaired mitochondrial dynamics and abnormal interaction of amyloid beta with mitochondrial protein Drp1 in neurons from patients with Alzheimer’s disease: Implications for neuronal damage. Hum. Mol. Genet. 2011, 20, 2495–2509.
  48. Wang, X.; Su, B.O.; Lee, H.G.; Li, X.; Perry, G.; Smith, M.A.; Zhu, X. Impaired balance of mitochondrial fission and fusion in Alzheimer’s disease. J. Neurosci. 2009, 29, 9090–9103.
  49. Beck, J.S.; Mufson, E.J.; Counts, S.E. Evidence for mitochondrial UPR gene activation in familial and sporadic Alzheimer’s Disease. Curr. Alzheimer Res. 2016, 13, 610–614.
  50. Sorrentino, V.; Romani, M.; Mouchiroud, L.; Beck, J.S.; Zhang, H.; D’Amico, D.; Moullan, N.; Potenza, F.; Schmid, W.A.; Rietsch, S.; et al. Enhancing mitochondrial proteostasis reduces amyloid-beta proteotoxicity. Nature 2017, 552, 187–193.
  51. Pérez, M.J.; Ivanyuk, D.; Panagiotakopoulou, V.; Di Napoli, G.; Kalb, S.; Brunetti, D.; Al-Shaana, R.; Kaeser, S.A.; Fraschka, S.A.-K.; Jucker, M.; et al. Loss of function of the mitochondrial peptidase PITRM1 induces proteotoxic stress and Alzheimer’s disease-like pathology in human cerebral organoids. Mol. Psych. 2021, 26, 5733–5750.
  52. Shen, Y.; Ding, M.; Xie, Z.; Liu, X.; Yang, H.; Jin, S.; Xu, S.; Zhu, Z.; Wang, Y.; Wang, D.; et al. Activation of Mitochondrial Unfolded Protein Response in SHSY5Y Expressing APP Cells and APP/PS1 Mice. Front. Cell. Neurosci. 2019, 13, 568.
  53. Weidling, I.W.; Swerdlow, R.H. Mitochondria in Alzheimer’s disease and their potential role in Alzheimer’s proteostasis. Exp. Neurol. 2020, 330, 113321.
  54. Landles, C.; Bates, G.P. Huntingtin and the molecular pathogenesis of Huntington’s disease: Fourth in molecular medicine review series. EMBO Rep. 2004, 5, 958–963.
  55. MacDonald, M.E.; Ambrose, C.M.; Duyao, M.P.; Myers, R.H.; Lin, C.; Srinidhi, L.; Barnes, G.; Taylor, S.A.; James, M.; Groot, N.; et al. A novel gene containing a trinucleotide repeat that is expanded and unstable on Huntington’s disease chromosomes. The Huntington’s Disease Collaborative Research Group. Cell 1993, 72, 971–983.
  56. Costa, V.; Scorrano, L. Shaping the role of mitochondria in the pathogenesis of Huntington’s disease. EMBO J. 2012, 31, 1853–1864.
  57. Bossy-Wetzel, E.; Petrilli, A.; Knott, A.B. Mutant huntingtin and mitochondrial dysfunction. Trends Neurosci. 2008, 31, 609–616.
  58. Pickrell, A.M.; Fukui, H.; Wang, X.; Pinto, M.; Moraes, C.T. The Striatum Is Highly Susceptible to Mitochondrial Oxidative Phosphorylation Dysfunctions. J. Neurosci. 2011, 31, 9895–9904.
  59. Fu, Z.; Liu, F.; Liu, C.; Jin, B.; Jiang, Y.; Tang, M.; Qi, X.; Guo, X. Mutant huntingtin inhibits the mitochondrial unfolded protein response by impairing ABCB10 mRNA stability. Biochim. Biophys. Acta—Mol. Basis Dis. 2019, 1865, 1428–1435.
  60. Yano, M. ABCB10 depletion reduces unfolded protein response in mitochondria. Biochem. Biophys. Res. Commun. 2017, 486, 465–469.
  61. Zhao, Q.; Wang, J.; Levichkin, I.V.; Stasinopoulos, S.; Ryan, M.; Hoogenraad, N.J. A mitochondrial specific stress response in mammalian cells. EMBO J. 2002, 21, 4411–4419.
  62. Rosen, D.R.; Siddique, T.; Patterson, D.; Figlewicz, D.A.; Sapp, P.; Hentati, A.; Donaldson, D.; Goto, J.; O’Regan, J.P.; Deng, H.-X.; et al. Mutations in Cu/Zn superoxide dismutase gene are associated with familial amyotrophic lateral sclerosis. Nature 1993, 362, 59–62.
  63. Renton, A.E.; Chio, A.; Traynor, B.J. State of play in amyotrophic lateral sclerosis genetics. Nat. Neurosci. 2014, 17, 17–23.
  64. Jawaid, A.; Khan, R.; Polymenidou, M.; Schulz, P.E. Disease-modifying effects of metabolic perturbations in ALS/FTLD. Mol. Neurodegener. 2018, 13, 63.
  65. Smith, E.F.; Shaw, P.J.; De Vos, K.J. The role of mitochondria in amyotrophic lateral sclerosis. Neurosci. Lett. 2019, 710, 132933.
  66. Prasad, A.; Bharathi, V.; Sivalingam, V.; Girdhar, A.; Patel, B.K. Molecular Mechanisms of TDP-43 Misfolding and Pathology in Amyotrophic Lateral Sclerosis. Front. Mol. Neurosci. 2019, 12, 25.
  67. Scotter, E.L.; Chen, H.J.; Shaw, C.E. TDP-43 Proteinopathy and ALS: Insights into disease mechanisms and therapeutic targets. Neurotherapeutics 2015, 12, 352–363.
  68. Wang, P.; Deng, J.; Dong, J.; Liu, J.; Bigio, E.H.; Mesulam, M.; Wang, T.; Sun, L.; Wang, L.; Lee, A.Y.-L.; et al. TDP-43 induces mitochondrial damage and activates the mitochondrial unfolded protein response. PLoS Genet. 2019, 15, e1007947.
  69. Zhou, Q.; Zhu, L.; Qiu, W.; Liu, Y.; Yang, F.; Chen, W.; Xu, R. Nicotinamide riboside enhances mitochondrial proteostasis and adult neurogenesis through activation of mitochondrial unfolded protein response signaling in the brain of ALS SOD1(G93A) mice. Int. J. Biol. Sci. 2020, 16, 284–297.
  70. Riar, A.K.; Burstein, S.R.; Palomo, G.M.; Arreguin, A.; Manfredi, G.; Germain, D. Sex specific activation of the ERalpha axis of the mitochondrial UPR (UPRmt) in the G93A-SOD1 mouse model of familial ALS. Hum. Mol. Genet. 2017, 26, 1318–1327.
  71. Gomez, M.; Germain, D. Cross talk between SOD1 and the mitochondrial UPR in cancer and neurodegeneration. Mol. Cell. Neurosci. 2019, 98, 12–18.
  72. Sciarretta, S.; Maejima, Y.; Zablocki, D.; Sadoshima, J. The Role of Autophagy in the Heart. Annu. Rev. Physiol. 2018, 80, 1–26.
  73. Nandi, S.S.; Katsurada, K.; Sharma, N.M.; Mahata, S.K.; Patel, K.P. Abstract 15288: Mitochondrial Injury in Cardiomyopathy of Neurogenic Hypertension: Role of MiR-18a-5p/HIF-1a Axis. Circulation 2020, 142 (Suppl. 3), A15288.
  74. Wrobel, L.; Topf, U.; Bragoszewski, P.; Wiese, S.; Sztolsztener, M.E.; Oeljeklaus, S.; Varabyova, A.; Lirski, M.; Chroscicki, P.; Mroczek, S.; et al. Mistargeted mitochondrial proteins activate a proteostatic response in the cytosol. Nature 2015, 524, 485–488.
  75. García-Díaz, L.; Coserria, F.; Antiñolo, G. Hypertrophic Cardiomyopathy due to Mitochondrial Disease: Prenatal Diagnosis, Management, and Outcome. Case Rep. Obstet. Gynecol. 2013, 2013, 472356.
  76. Smyrnias, I.; Gray, S.P.; Okonko, D.O.; Sawyer, G.; Zoccarato, A.; Catibog, N.; López, B.; Gonzalez, A.; Ravassa, S.; Díez, J.; et al. Cardioprotective Effect of the Mitochondrial Unfolded Protein Response During Chronic Pressure Overload. J. Am. Coll. Cardiol. 2019, 73, 1795–1806.
  77. Wang, Y.; Jasper, H.; Toan, S.; Muid, D.; Chang, X.; Zhou, H. Mitophagy coordinates the mitochondrial unfolded protein response to attenuate inflammation-mediated myocardial injury. Redox Biol. 2021, 45, 102049.
  78. Houtkooper, R.; Mouchiroud, L.; Ryu, D.; Moullan, N.; Katsyuba, E.; Knott, G.W.; Williams, R.W.; Auwerx, J. Mitonuclear protein imbalance as a conserved longevity mechanism. Nature 2013, 497, 451–457.
  79. Durieux, J.; Wolff, S.; Dillin, A. The Cell-Non-Autonomous Nature of Electron Transport Chain-Mediated Longevity. Cell 2011, 144, 79–91.
  80. Yokoyama, K.; Fukumoto, K.; Murakami, T.; Harada, S.I.; Hosono, R.; Wadhwa, R.; Mitsui, Y.; Ohkuma, S. Extended longevity of Caenorhabditis elegans by knocking in extra copies of hsp70F, a homolog of mot-2 (mortalin)/mthsp70/Grp75. FEBS Lett. 2002, 516, 53–57.
  81. Flurkey, K.; Papaconstantinou, J.; Miller, R.A.; Harrison, D.E. Lifespan extension and delayed immune and collagen aging in mutant mice with defects in growth hormone production. Proc. Natl. Acad. Sci. USA 2001, 98, 6736–6741.
  82. Ozkurede, U.; Miller, R.A. Improved mitochondrial stress response in long-lived Snell dwarf mice. Aging Cell 2019, 18, e13030.
  83. Wei, Y.; Zhang, Y.J.; Cai, Y.; Xu, M.H. The role of mitochondria in mTOR-regulated longevity. Biol. Rev. Camb. Philos. Soc. 2015, 90, 167–181.
  84. Jensen, M.B.; Jasper, H. Mitochondrial Proteostasis in the Control of Aging and Longevity. Cell Metab. 2014, 20, 214–225.
  85. Haynes, C.M.; Yang, Y.; Blais, S.P.; Neubert, T.A.; Ron, D. The Matrix Peptide Exporter HAF-1 Signals a Mitochondrial UPR by Activating the Transcription Factor ZC376.7 in C. elegans. Mol. Cell 2010, 37, 529–540.
  86. Ristow, M.; Schmeisser, S. Extending life span by increasing oxidative stress. Free Radic. Biol. Med. 2011, 51, 327–336.
  87. Mouchiroud, L.; Houtkooper, R.H.; Moullan, N.; Katsyuba, E.; Ryu, D.; Cantó, C.; Mottis, A.; Jo, Y.-S.; Viswantahan, M.; Schoonjans, K.; et al. The NAD(+)/sirtuin pathway modulates longevity through activation of mitochondrial UPR and FOXO signaling. Cell 2013, 154, 430–441.
  88. Schmeisser, K.; Mansfeld, J.; Kuhlow, D.; Weimer, S.; Priebe, S.; Heiland, I.; Birringer, M.; Groth, M.; Segref, A.; Kanfi, Y.; et al. Role of sirtuins in lifespan regulation is linked to methylation of nicotinamide. Nat. Chem. Biol. 2013, 9, 693–700.
  89. Hsu, A.-L.; Murphy, C.T.; Kenyon, C. Regulation of aging and age-related disease by DAF-16 and heat-shock factor. Science 2003, 300, 1142–1145.
  90. Yang, W.; Hekimi, S. Two modes of mitochondrial dysfunction lead independently to lifespan extension in Caenorhabditis elegans. Aging Cell 2010, 9, 433–447.
  91. Zhang, H.; Ryu, D.; Wu, Y.; Gariani, K.; Wang, X.; Luan, P.; D’Amico, D.; Ropelle, E.R.; Lutolf, M.P.; Aebersold, R.; et al. NAD+ repletion improves mitochondrial and stem cell function and enhances life span in mice. Science 2016, 352, 1436–1443.
  92. Behrends, C.; Langer, C.A.; Boteva, R.; Böttcher, U.M.; Stemp, M.J.; Schaffar, G.; Rao, B.V.; Giese, A.; Kretzschmar, H.; Siegers, K.; et al. Chaperonin TRiC Promotes the Assembly of polyQ Expansion Proteins into Nontoxic Oligomers. Mol. Cell 2006, 23, 887–897.
More
Information
Subjects: Cell Biology
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , , , , , , ,
View Times: 405
Revisions: 2 times (View History)
Update Date: 22 Jul 2022
1000/1000
Video Production Service