Next Article in Journal
Zebrafish Models of Autosomal Dominant Ataxias
Next Article in Special Issue
Rapid Identification of New Biomarkers for the Classification of GM1 Type 2 Gangliosidosis Using an Unbiased 1H NMR-Linked Metabolomics Strategy
Previous Article in Journal
A LAT-Based Signaling Complex in the Immunological Synapse as Determined with Live Cell Imaging Is Less Stable in T Cells with Regulatory Capability
Previous Article in Special Issue
Optimization of Enzyme Essays to Enhance Reliability of Activity Measurements in Leukocyte Lysates for the Diagnosis of Metachromatic Leukodystrophy and Gangliosidoses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Potential Treatment of Lysosomal Storage Disease through Modulation of the Mitochondrial—Lysosomal Axis

1
Division of Life Sciences, College of Life Sciences and Bioengineering, Incheon National University, Incheon 22012, Korea
2
The Future Life & Society Research Center, Chonnam National University, Gwangju 61186, Korea
*
Authors to whom correspondence should be addressed.
Cells 2021, 10(2), 420; https://doi.org/10.3390/cells10020420
Submission received: 26 January 2021 / Revised: 13 February 2021 / Accepted: 14 February 2021 / Published: 17 February 2021
(This article belongs to the Special Issue Lysosomal Storage Disorders)

Abstract

:
Lysosomal storage disease (LSD) is an inherited metabolic disorder caused by enzyme deficiency in lysosomes. Some treatments for LSD can slow progression, but there are no effective treatments to restore the pathological phenotype to normal levels. Lysosomes and mitochondria interact with each other, and this crosstalk plays a role in the maintenance of cellular homeostasis. Deficiency of lysosome enzymes in LSD impairs the turnover of mitochondrial defects, leading to deterioration of the mitochondrial respiratory chain (MRC). Cells with MRC impairment are associated with reduced lysosomal calcium homeostasis, resulting in impaired autophagic and endolysosomal function. This malicious feedback loop between lysosomes and mitochondria exacerbates LSD. In this review, we assess the interactions between mitochondria and lysosomes and propose the mitochondrial–lysosomal axis as a research target to treat LSD. The importance of the mitochondrial–lysosomal axis has been systematically characterized in several studies, suggesting that proper regulation of this axis represents an important investigative guide for the development of therapeutics for LSD. Therefore, studying the mitochondrial–lysosomal axis will not only add knowledge of the essential physiological processes of LSD, but also provide new strategies for treatment of LSD.

1. Introduction

Lysosomes are membrane-bound organelles that contain a variety of digestive enzymes [1]. Lysosomes are involved in several cellular processes, including macromolecular catabolism and recycling [1]. Lysosomal storage disease (LSD) is a disorder caused by a deficiency of enzymes in lysosomes, leading to the accumulation of undigested or partially digested macromolecules [2,3]. LSD includes 50 rare monogenic disorders, classified according to the specific genetic defects and biochemical properties of the stored macromolecules [4,5,6,7,8,9]. LSD affects all parts of the body, including the skeleton, skin, heart, and central nervous system [10]. Recently, enzyme replacement therapy has been suggested as a treatment for LSD as it can alleviate symptoms [11,12,13]. However, this treatment has not been effective in restoring pathologic phenotypes to normal levels [11,12,13]. Although other clinical trials are ongoing for possible treatments for some of these diseases, there are currently no approved treatments for LSD.
Lysosomes and mitochondria have classical functions, namely as recycling bins and energy factories, respectively [14,15]. Mounting evidence supports functional interconnection and suggests it as a signaling platform to maintain cellular homeostasis [16]. Recently, a functional role of the mitochondrial–lysosomal axis in senescence alleviation has been proposed [14,17]. Proper control of this axis is essential for senescence alleviation, as evidenced by experiments showing that activation of this axis, by mitochondrial functional recovery, restored the senescent phenotypes [14,17]. Furthermore, the importance of this axis in senescence is emphasized by the fact that activation of this axis by lysosomal acidification has endowed an immortal germ lineage in Caenorhabditis elegans [18]. Thus, the modulation of this axis has been proposed as an important research guideline for the development of therapeutics for senescence [14,17].
The purpose of this review is to summarize LSD-related lysosomal/mitochondrial dysfunction and speculate on the crosstalk between these organelles. An extensive literature search in PubMed (https://pubmed.ncbi.nlm.nih.gov/) (accessed on 15 October 2020) was performed using search terms such as LSD-related lysosomal dysfunction, LSD-related mitochondrial dysfunction, mitochondrial–lysosomal crosstalk, and mitochondrial–lysosomal axis. In order to process and analyze the collected data and results, we used a systematic review strategy known as the gold standard for review writing in medical fields [19]. Based on the previous and recent findings that have been processed and analyzed (Supplementary Figure S1), we will highlight the role of the mitochondrial–lysosomal axis in LSD and propose a crucial role for this axis as a regulator of LSD treatment.

2. Lysosomal Dysfunction in LSD

Lysosomes are membrane-bound organelles containing various lytic enzymes within the acidic vacuolar compartment [20,21]. They incorporate new hydrolytic enzymes by fusion with vesicles pinched off from the Golgi apparatus and break down a variety of substrates ranging from intracellular macromolecules to impaired organelles (Figure 1A; blue dot indicates functional lysosomal enzymes). Lysosomes fuse with intermediate vesicles, termed autophagosomes, where cargos are either sequestered in bulk (e.g., cytosols with all the contents in that area) or selectively (e.g., mitochondria, endoplasmic reticulum or ribosomes) [22]. Subsequent fusion of autophagosomes with lysosomes creates autolysosomes, where lysosomal enzymes degrade the autophagosome-delivered substrates (Figure 1A). Extracellular substances are internalized and delivered from the plasma membrane to the endosome (Figure 1A). Lysosomes fuse with endosomes to form highly dynamic membrane structures called endolysosomes, in which endocytosed substrates are degraded and recycled (Figure 1A). The digestion of autophagocytosed or endocytosed substrates occurs in an acidic environment of the lysosomal lumen (pH 4.5 to 5.0), which is maintained by a vacuolar ATPase (V-ATPase) proton pump [23,24]. V-ATPase uses the energy of ATP hydrolysis to move protons (H+) from the cytoplasm to the lysosome lumen against a concentration gradient (Figure 1A). In addition, the digestion of substrates occurs at an appropriate level of calcium ions, which are maintained by mammalian mucolipin TRP (Transient Receptor Potential) channel subfamily (TRPML1) transporters in the lysosomal membrane [25]. Lysosomal Ca2+ homeostasis plays an important role in the regulation of lysosomal function, including vesicular trafficking and fusion processes [26].
LSD is a metabolic disorder caused by mutations in genes that encode lysosomal enzymes such as lysosomal glycosidases, proteases, enzyme modifiers or activators [27,28]. Mutations alter the functional activity of lysosomal enzymes, leading to abnormal degradation of substrates and accumulation of various substrates inside the lysosome (i.e., “storage”) (Figure 1B; red dot indicates dysfunctional lysosomal enzymes). These events exacerbate the activity of lysosomal enzymes that are not genetically deficient (Figure 1B). Lysosomes loaded with undigested substances exhibit defective fusion with autophagosomes, leading to the accumulation of unfused autophagosomes (Figure 1B). For example, LSD cells have an increased number of organelles containing autophagosome markers, but only a few of them display lysosomal markers, suggesting defective fusion between lysosomes and autophagosomes [29]. Furthermore, lysosomes of Niemann–Pick type A/C disease, a type of LSD, exhibit alterations in endolysosome trafficking characterized by the accumulation of extracellular substances and specific lipid species in lysosomes [30]. The altered lysosomal pH represents another characteristic of LSD [31,32]. In LSD, lysosomes loaded with undigested lipids with free amine groups in cationic lipids cause severe lysosomal and cellular dysfunction, including abnormal alkalization of lysosomal pH [31]. The altered lysosomal pH subsequently deteriorates the activity of various lysosomal enzymes (Figure 1B). In addition, dysregulation of lysosomal Ca2+ homeostasis constitutes another feature of LSD (Figure 1B). For example, one type of LSD, Mucolipidosis type IV (MLIV), is caused by a loss-of-function mutation in TRPML1 associated with lysosome-mediated Ca2+ efflux [33]. The decreased Ca2+ release via TRPML1 disrupts endocytosis and vesicular fusion, which limits the access of lipids to lysosomal lipases, along with the accumulation of undegraded lipids in MLIV [33].

3. Mitochondrial Dysfunction in LSD

Mitochondria are double-membrane organelles that produce most of the cellular ATP, which is used as a source of a cell’s biochemical reaction [34,35]. These highly dynamic organelles can be processed through successive fusion and fission cycles (Figure 2A). Cytoplasmic Ca2+ modulates GTPase activity, which regulates the balance of mitochondrial fusion/fission events (Figure 2A). The optimal Ca2+ concentration in the cytoplasm is primarily controlled by lysosomal Ca2+ efflux through TRPML1 and Ca2+ release through the endoplasmic reticulum (ER) (Figure 2A) [36]. Although mitochondrial fusion contributes to a relatively homogeneous network by mixing the contents of partially damaged mitochondria, mitochondrial fission is required to generate new mitochondria by eliminating damaged mitochondria through autophagy [37]. Functional lysosomes are essential for autophagy-mediated removal of defective mitochondria (mitophagy; Figure 2A). Specifically, mitophagy promotes turnover of mitochondria and prevents the accumulation of dysfunctional mitochondria [38].
Abnormal mitochondrial morphology is a pathogenic feature of LSD [39,40]. Dysregulation of lysosomal Ca2+ homeostasis in LSD interferes with the balance of mitochondrial fission and fusion cycles, leading to mitochondrial morphology changes, namely mitochondrial fragmentation or elongation (Figure 2B). These abnormalities are associated with degenerative mitochondrial functions, such as qualitative changes in mitochondrial membrane potential (ΔΨm) and MRC function [41,42]. Support for this phenomenon is evident from observations showing that a mouse model of LSD, produced by knocking out sulfatase-modifying factor 1, exhibits mitochondrial fragmentation accompanying impaired ATP production [43]. Moreover, fragmented mitochondria with reduced ΔΨm are present in neurons and astrocytes cultured from another LSD mouse model, GM1 gangliosidosis [44]. Furthermore, the elongated mitochondrial morphology as a characteristic in LSD can be inferred by evidence showing the accumulation of swollen and elongated mitochondria with disorganized crista in LSD cells [45]. The elongated mitochondria also exhibit reduced MRC function concomitant with decreased ATP production. Collectively, these observations show that abnormal mitochondrial morphology, primarily attributed to the perturbation of Ca2+ homeostasis, constitutes one of the salient features observed in LSD.
Accumulation of dysfunctional mitochondria is a surrogate marker for abnormal mitochondrial morphology in LSD. The accumulation of impaired mitochondria is primarily because of lysosomal deficiency in LSD [42]. In particular, lysosomal deficiency deteriorates Ca2+ buffering, consequently inhibiting the sequential stages of pre-autophagosome formation and autophagosome maturation (Figure 2B). Even after lysosomal fusion with autophagosomes, inactive hydrolytic enzymes in lysosomes are unable to degrade dysfunctional mitochondria (Figure 2B). For example, in LSD because of alpha-glucosidase deficiency, dysfunctional mitochondria are sequestered in autophagy vesicles, but not digested by lysosomes, suggesting inefficient lysosomal-autophagy-mediated degradation (Figure 2B) [46]. Support for this finding is evident in previous results showing that the accumulation of mitochondria with various abnormalities was observed in LSD, including MLII, MLIII [47], MLIV [48], GM1 gangliosidosis [44,49], or Batten disease [50]. Dysfunctional mitochondria are not efficiently eliminated and constitute a major source of reactive oxygen species (ROS) production. Excessive ROS causes severe damage to lysosomes, exacerbating the accumulation of damaged mitochondria [51]. The accumulation of dysfunctional mitochondria is a well-known feature of LSD and leads to deleterious consequences, including altered cellular homeostasis and tissue degeneration [52].

4. Importance of the Mitochondrial–Lysosomal Axis in LSD

Lysosomes and mitochondria interact with each other, and this crosstalk can act as a central hub at various levels [14,16]. They constantly inform each other of their functional status through retrograde or anterograde signals [53]. These organelles are recognized as signaling platforms that regulate many key elements of cellular physiology and are more than simple units, with interconnections through the mitochondrial–lysosomal axis [16].
The effect of lysosomal function on mitochondria is systematically characterized, indicating that the regulatory mechanisms of lysosomes control mitochondrial function. As mentioned in Section 3, lysosomal defects in LSD impair the mechanism by which the damaged mitochondria are removed (Figure 3A). Thus, damaged mitochondria accumulate concomitant declines in mitochondrial function, such as decreased ATP production and increased ROS generation (Figure 3A). In most cases, lysosomes regulate mitochondrial function without physical contact, but sometimes lysosomes control mitochondrial function through interorganelle contacts [54,55]. For example, Ca2+ transport is mediated by the lysosomal channel TRPML1 in lysosome–mitochondrial contact [56]. Thus, a deficiency of TRPML1 function in MLIV results in a decrease in Ca2+ concentration at the lysosomal–mitochondrial interface [56]. In turn, the decrease in Ca2+ concentration reduces contact-dependent mitochondrial calcium absorption by the mitochondrial calcium uniporter (MCU), impairing calcium-dependent mitochondrial function, including oxidative phosphorylation and ATP production [33,57,58]. Therefore, these findings support a potential role for lysosome in the regulation of mitochondria, in a contact-independent or dependent manner.
The reciprocal crosstalk between mitochondria and lysosomes suggests that the regulatory mechanisms of mitochondrial function are essential for the proper functioning of lysosomes. For example, a mouse model with a genetic deletion of mitochondrial transcription factor A (Tfam) was generated, and the effects of mitochondria on lysosomes were evaluated [59]. Tfam encodes a mitochondrial transcription factor that plays an essential role in the transcription and replication of the mitochondrial genome [60,61]. Tfam-deficient cells show abnormal levels of mitochondrial DNA expression concomitant with reduced MRC function and altered mitochondrial metabolism (Figure 3B). Cells with reduced MRC function are associated with reduced lysosomal Ca2+ efflux, resulting in impaired endolysosomal function with an abnormal accumulation of lipid species, including sphingomyelin (Figure 3B). Extending the relevance of these findings, the deficiency of the Tfam deteriorates mitochondrial function, which was accompanied by disturbances in endolysosomal trafficking and autophagy function [59]. Thus, these findings suggest that there is an effect of mitochondria on the regulation of lysosomal function, which underlies the operation of the mitochondrial–lysosomal axis.
The functional interconnection between lysosomes and mitochondria is highlighted by the fact that lysosomal acidification is achieved by V-ATPases, which require energy in the form of ATP energies, which are provided by the functional mitochondria [62,63]. As noted in Section 2, V-ATPase contributes to maintaining an acidic environment in the lysosomal lumen, allowing most of the lysosomal enzymes to be active (Figure 3C). However, reduction of MRC function in LSD prevents mitochondria from supplying sufficient ATP to V-ATPase for proper functioning in the lysosome membrane [64,65]. This finding has been established in recent studies showing that inhibition of mitochondrial function after deletion of the mitochondrial proteins (AIF, OPA1 or PINK1) reduces the activity of MRC, thereby reducing ATP production [66]. Furthermore, in cells with reduced MRC function, lysosomal pH did not change even after suppressing V-ATPase activity with bafilomycin, but in normal cells, lysosomal pH changed after suppressing V-ATPase [66]. These results indicate that functional mitochondria are necessary for V-ATPase-mediated lysosomal acidification. Thus, the functional interdependence between these organelles suggests the role of the mitochondrial–lysosomal axis as a promising and potent target for LSD treatment.

5. Strategies to Activate the Mitochondrial–Lysosomal Axis in LSD

Mitochondria and lysosomes constantly interact with each other in a reciprocal manner. Functional interconnection via the mitochondrial–lysosomal axis suggests that the regulatory mechanisms of one organelle are essential for the proper functioning of another organelle [14,16]. Therefore, it is appropriate to assume that proper control of the mitochondrial–lysosomal axis might represent a new therapeutic frontier in LSD treatment.
Recent studies have demonstrated that the pathophysiologic symptoms of LSD could be restored to a normal state by activation of the mitochondrial–lysosomal axis [67,68,69]. Specifically, Gaucher’s disease (GD), caused by a defect in lysosome β-glucocerebrosidase (GCase), exhibits impaired electron transport in MRC coupled with increased mitochondrial ROS generation and decreased ATP production [70]. In accordance with this finding, GD cells have a significantly lower content of coenzyme Q10 (CoQ), which is known to enhance MRC activity by accepting one electron from complexes I/II and transferring the reducing equivalent to complex III (Figure 4A; Green CoQ indicates a low content of CoQ). Given that mitochondrial MRC deficiency plays a crucial role in the pathogenesis of GD, GD cells have been supplemented with CoQ [71]. CoQ supplementation results in an improvement in mitochondrial function, indicated by a marked increase in cellular ATP levels and a marked decrease in ROS generation (Figure 4A; Pink CoQ indicates a high content of CoQ). Restoration of mitochondrial function ameliorates GCase defects in GD, resulting in reduced glycosaminoglycan accumulation [70]. The significance of mitochondrial functional recovery on LSD treatment is further supported by experiments with NAD+, which acts as an electron carrier in MRC during oxidative phosphorylation [72]. Indeed, increasing intracellular levels of NAD+ led to marked improvement of the MRC defect [67,68,69]. Specifically, in mice lacking acid sphingomyelinase (ASM) activity, NAD+ treatment enhances ASM activity, reducing substrate accumulation in lysosomes [59]. These findings suggest that a strategy to activate the mitochondrial–lysosomal axis through restoration of mitochondrial function might be effective in treating LSD.
The mitochondrial–lysosomal axis is not unidirectional but functions bidirectionally to treat LSDs. For example, activation of autophagy with trehalose, an autophagy inducer, improves pathological symptoms in a mouse model of mucopolysaccharide IIIB, a type of LSD [73]. Specifically, trehalose treatment efficiently clears autophagy vacuoles from neurons, along with activation of the transcription factor EB (TFEB), which regulates lysosome biogenesis (Figure 4B). Furthermore, improvement of LSD symptoms by trehalose was blocked by Atg7 knockdown, indicating that autophagy activation is a prerequisite for treating neurological symptoms associated with LSD [73]. Similar to this observation, promoting autophagy maturation and subsequent exocytosis in LSD enables the rescue of pathological storage and restoration of normal cellular state [74]. These data suggest that activation of the mitochondrial–lysosomal axis through activation of autophagy might be an alternative strategy for LSD treatment.
The regulation of the mitochondrial–lysosomal axis via lysosomal acidification represents a promising potent strategy for LSD treatment. Almost all LSDs have shown defects in lysosomal acidification, which provides an optimal environment for the degradation of macromolecules delivered to the lysosome and facilitates autophagic flux [30,75]. Lysosomal acidification is regulated by V-ATPase, composed of a cytoplasmic V1 sector and a lysosomal membrane-bound V0 sector. The reversible assembly of the V1/V0 sectors induces a functionally active V-ATPase to maintain lysosomal acidification [21,76]. A recent study revealed that V1/V0 assembly of V-ATPase is regulated by the inhibition of Ataxia telangiectasia mutated (ATM) gene [77]. Specifically, attenuation of ATM activity promotes V1/V0 assembly in V-ATPase accompanied by re-acidification of lysosomes (Figure 4C). Lysosomal re-acidification restores the activity of lysosomal enzymes and increases autophagic flux, which in turn accelerates the elimination of dysfunctional mitochondria with restoration of mitochondrial function [77]. Thus, activation of the mitochondrial–lysosomal axis through lysosomal re-acidification may enable an optimal environment for lysosomal enzymes, providing therapeutic benefits in situations of lysosomal dysfunction.

6. Conclusions and Perspectives

Functional interconnection between lysosomes and mitochondria has been reviewed in several review articles. One review specifically addressed the coordination between mitochondria and lysosome with a focus on cellular metabolism and signaling [53]. The importance of reciprocal crosstalk was discussed but was limited to its role in neurodegeneration [53]. This limitation of relevance was improved by another review addressing the potential role of mitochondrial dysfunction in the pathophysiology of LSD [42]. This review suggested that mitochondrial damage can lead to lysosomal dysfunction, supporting a common signaling pathway and crosstalk between the two organelles [42]. Although the causes and consequences of mitochondrial damage have been extensively described, no specific strategy to induce mitochondrial function recovery has been proposed [42]. Most recently, a review paper has been published focusing on mitochondrial dysfunction as an important contributing factor in the pathophysiology of LSD [78]. This review explained that dysfunctional mitochondria affect the function of lysosomes by generating ROS or depriving lysosomes of ATP [78]. Thus, a putative mechanism capable of causing mitochondrial dysfunction in LSD has been proposed as a therapeutically effective target in patients with LSD [78]. However, in LSD, many cellular pathways are impaired, so it remains unclear whether treatment strategies aimed at improving only mitochondrial dysfunction will be effective in the treatment of these disorders.
In this review, we summarized LSD-related lysosomal/mitochondrial dysfunction and investigated the crosstalk between lysosomes and mitochondria. Functional interconnections between lysosomes and mitochondria suggest that the regulatory mechanisms of one organelle are required for the proper functioning of another organelle. Thus, we proposed the mitochondrial–lysosomal axis as a potential therapeutic target for treating LSD (Figure 5). Regulation of this axis is bidirectional, either through restoration of mitochondrial function or restoration of lysosome function. Proper regulation of this axis, which is not dependent on one approach, may rescue the pathological symptoms of LSD, where many cellular pathways are impaired (Figure 5). A deeper understanding of the molecular mechanisms supporting the role of this axis in the onset and progression of LSD will pave the way for effective treatment of patients with LSD.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4409/10/2/420/s1, Figure S1: Information on the year of publication of research papers referenced in this review. Forty-seven percent of the references have been published since 2015, indicating that this review article is up to date.

Author Contributions

M.U.K., Y.H.L., J.W.K., S.Y.H., J.T.P. and S.C.P. wrote and edited the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT & Future Planning (NRF-2020 R1A2C2009439) and the Research Assistance Program (2020) in the Incheon National University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Luzio, J.P.; Pryor, P.R.; Bright, N.A. Lysosomes: Fusion and function. Nat. Rev. Mol. Cell Biol. 2007, 8, 622–632. [Google Scholar] [CrossRef]
  2. Sun, A. Lysosomal storage disease overview. Ann. Transl. Med. 2018, 6, 476. [Google Scholar] [CrossRef]
  3. Platt, F.M.; d’Azzo, A.; Davidson, B.L.; Neufeld, E.F.; Tifft, C.J. Lysosomal storage diseases. Nat. Rev. Dis. Primers 2018, 4, 27. [Google Scholar] [CrossRef] [PubMed]
  4. Chapel, A.; Kieffer-Jaquinod, S.; Sagné, C.; Verdon, Q.; Ivaldi, C.; Mellal, M.; Thirion, J.; Jadot, M.; Bruley, C.; Garin, J.; et al. An extended proteome map of the lysosomal membrane reveals novel potential transporters. Mol. Cell. Proteom. MCP 2013, 12, 1572–1588. [Google Scholar] [CrossRef] [Green Version]
  5. Di Fruscio, G.; Schulz, A.; De Cegli, R.; Savarese, M.; Mutarelli, M.; Parenti, G.; Banfi, S.; Braulke, T.; Nigro, V.; Ballabio, A. Lysoplex: An efficient toolkit to detect DNA sequence variations in the autophagy-lysosomal pathway. Autophagy 2015, 11, 928–938. [Google Scholar] [CrossRef]
  6. Palmieri, M.; Impey, S.; Kang, H.; di Ronza, A.; Pelz, C.; Sardiello, M.; Ballabio, A. Characterization of the CLEAR network reveals an integrated control of cellular clearance pathways. Hum. Mol. Genet. 2011, 20, 3852–3866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Schröder, B.; Wrocklage, C.; Pan, C.; Jäger, R.; Kösters, B.; Schäfer, H.; Elsässer, H.P.; Mann, M.; Hasilik, A. Integral and associated lysosomal membrane proteins. Traffic (Cph. Den.) 2007, 8, 1676–1686. [Google Scholar] [CrossRef]
  8. Sleat, D.E.; Ding, L.; Wang, S.; Zhao, C.; Wang, Y.; Xin, W.; Zheng, H.; Moore, D.F.; Sims, K.B.; Lobel, P. Mass spectrometry-based protein profiling to determine the cause of lysosomal storage diseases of unknown etiology. Mol. Cell. Proteom. MCP 2009, 8, 1708–1718. [Google Scholar] [CrossRef] [Green Version]
  9. Szklarczyk, D.; Franceschini, A.; Wyder, S.; Forslund, K.; Heller, D.; Huerta-Cepas, J.; Simonovic, M.; Roth, A.; Santos, A.; Tsafou, K.P.; et al. STRING v10: Protein-protein interaction networks, integrated over the tree of life. Nucleic Acids Res. 2015, 43, D447–D452. [Google Scholar] [CrossRef]
  10. Ferreira, C.R.; Gahl, W.A. Lysosomal storage diseases. Transl. Sci. Rare Dis. 2017, 2, 1–71. [Google Scholar] [CrossRef] [Green Version]
  11. Li, M. Enzyme Replacement Therapy: A Review and Its Role in Treating Lysosomal Storage Diseases. Pediatr. Ann. 2018, 47, e191–e197. [Google Scholar] [CrossRef] [PubMed]
  12. Bruni, S.; Loschi, L.; Incerti, C.; Gabrielli, O.; Coppa, G.V. Update on treatment of lysosomal storage diseases. Acta Myol. 2007, 26, 87–92. [Google Scholar] [PubMed]
  13. Parenti, G.; Pignata, C.; Vajro, P.; Salerno, M. New strategies for the treatment of lysosomal storage diseases (Review). Int. J. Mol. Med. 2013, 31, 11–20. [Google Scholar] [CrossRef] [PubMed]
  14. Park, J.T.; Lee, Y.-S.; Cho, K.A.; Park, S.C. Adjustment of the lysosomal-mitochondrial axis for control of cellular senescence. Ageing Res. Rev. 2018, 47, 176–182. [Google Scholar] [CrossRef]
  15. Todkar, K.; Ilamathi, H.S.; Germain, M. Mitochondria and Lysosomes: Discovering Bonds. Front. Cell Dev. Biol. 2017, 5, 106. [Google Scholar] [CrossRef] [Green Version]
  16. Ohya, Y.; Umemoto, N.; Tanida, I.; Ohta, A.; Iida, H.; Anraku, Y. Calcium-sensitive cls mutants of Saccharomyces cerevisiae showing a Pet- phenotype are ascribable to defects of vacuolar membrane H+-ATPase activity. J. Biol. Chem. 1991, 266, 13971–13977. [Google Scholar] [CrossRef]
  17. Brunk, U.T.; Terman, A. The mitochondrial-lysosomal axis theory of aging: Accumulation of damaged mitochondria as a result of imperfect autophagocytosis. Eur. J. Biochem. 2002, 269, 1996–2002. [Google Scholar] [CrossRef] [PubMed]
  18. Bohnert, K.A.; Kenyon, C. A lysosomal switch triggers proteostasis renewal in the immortal C. elegans germ lineage. Nature 2017, 551, 629–633. [Google Scholar] [CrossRef] [Green Version]
  19. Liberati, A.; Altman, D.G.; Tetzlaff, J.; Mulrow, C.; Gøtzsche, P.C.; Ioannidis, J.P.A.; Clarke, M.; Devereaux, P.J.; Kleijnen, J.; Moher, D. The PRISMA statement for reporting systematic reviews and meta-analyses of studies that evaluate healthcare interventions: Explanation and elaboration. BMJ 2009, 339, b2700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Xu, H.; Ren, D. Lysosomal physiology. Annu. Rev. Physiol. 2015, 77, 57–80. [Google Scholar] [CrossRef] [Green Version]
  21. Bonam, S.R.; Wang, F.; Muller, S. Lysosomes as a therapeutic target. Nat. Rev. Drug Discov. 2019, 18, 923–948. [Google Scholar] [CrossRef] [Green Version]
  22. Singh, R.; Cuervo, A.M. Autophagy in the cellular energetic balance. Cell Metab. 2011, 13, 495–504. [Google Scholar] [CrossRef] [Green Version]
  23. Forgac, M. Vacuolar ATPases: Rotary proton pumps in physiology and pathophysiology. Nat. Rev. Mol. Cell Biol. 2007, 8, 917–929. [Google Scholar] [CrossRef]
  24. Nishi, T.; Forgac, M. The vacuolar H+-ATPases—Nature’s most versatile proton pumps. Nat. Rev. Mol. Cell Biol. 2002, 3, 94–103. [Google Scholar] [CrossRef]
  25. Lloyd-Evans, E.; Platt, F.M. Lysosomal Ca2+ homeostasis: Role in pathogenesis of lysosomal storage diseases. Cell Calcium 2011, 50, 200–205. [Google Scholar] [CrossRef] [PubMed]
  26. Lloyd-Evans, E.; Waller-Evans, H.; Peterneva, K.; Platt, F.M. Endolysosomal calcium regulation and disease. Biochem. Soc. Trans. 2010, 38, 1458–1464. [Google Scholar] [CrossRef] [Green Version]
  27. Ozand, P.T.; Al-Essa, M. Lysosomal Storage Diseases. In Textbook of Clinical Pediatrics; Elzouki, A.Y., Harfi, H.A., Nazer, H.M., Stapleton, F.B., Oh, W., Whitley, R.J., Eds.; Springer: Berlin/Heidelberg, Germany, 2012; pp. 515–555. [Google Scholar]
  28. Marques, A.R.A.; Saftig, P. Lysosomal storage disorders—Challenges, concepts and avenues for therapy: Beyond rare diseases. J. Cell Sci. 2019, 132, jcs221739. [Google Scholar] [CrossRef] [PubMed]
  29. Lieberman, A.P.; Puertollano, R.; Raben, N.; Slaugenhaupt, S.; Walkley, S.U.; Ballabio, A. Autophagy in lysosomal storage disorders. Autophagy 2012, 8, 719–730. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Futerman, A.H.; van Meer, G. The cell biology of lysosomal storage disorders. Nat. Rev. Mol. Cell Biol. 2004, 5, 554–565. [Google Scholar] [CrossRef]
  31. Folts, C.J.; Scott-Hewitt, N.; Pröschel, C.; Mayer-Pröschel, M.; Noble, M. Lysosomal Re-acidification Prevents Lysosphingolipid-Induced Lysosomal Impairment and Cellular Toxicity. PLoS Biol. 2016, 14, e1002583. [Google Scholar] [CrossRef] [Green Version]
  32. Parkinson-Lawrence, E.J.; Shandala, T.; Prodoehl, M.; Plew, R.; Borlace, G.N.; Brooks, D.A. Lysosomal Storage Disease: Revealing Lysosomal Function and Physiology. Physiology 2010, 25, 102–115. [Google Scholar] [CrossRef] [Green Version]
  33. Peng, W.; Wong, Y.C.; Krainc, D. Mitochondria-lysosome contacts regulate mitochondrial Ca2+ dynamics via lysosomal TRPML1. Proc. Natl. Acad. Sci. USA 2020, 117, 19266–19275. [Google Scholar] [CrossRef]
  34. Henze, K.; Martin, W. Evolutionary biology: Essence of mitochondria. Nature 2003, 426, 127–128. [Google Scholar] [CrossRef]
  35. Bernhardt, D.; Müller, M.; Reichert, A.S.; Osiewacz, H.D. Simultaneous impairment of mitochondrial fission and fusion reduces mitophagy and shortens replicative lifespan. Sci. Rep. 2015, 5, 7885. [Google Scholar] [CrossRef] [Green Version]
  36. Szabadkai, G.; Simoni, A.M.; Bianchi, K.; De Stefani, D.; Leo, S.; Wieckowski, M.R.; Rizzuto, R. Mitochondrial dynamics and Ca2+ signaling. Biochim. Biophys. Acta (BBA) Mol. Cell Res. 2006, 1763, 442–449. [Google Scholar] [CrossRef]
  37. Okamoto, K.; Kondo-Okamoto, N. Mitochondria and autophagy: Critical interplay between the two homeostats. BBA Gen Subj. 2012, 1820, 595–600. [Google Scholar] [CrossRef] [PubMed]
  38. Kiselyov, K.; Jennigs, J.J., Jr.; Rbaibi, Y.; Chu, C.T. Autophagy, mitochondria and cell death in lysosomal storage diseases. Autophagy 2007, 3, 259–262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Suzuki, K.; Yamaguchi, A.; Yamanaka, S.; Kanzaki, S.; Kawashima, M.; Togo, T.; Katsuse, O.; Koumitsu, N.; Aoki, N.; Iseki, E.; et al. Accumulated α-synuclein affects the progression of GM2 gangliosidoses. Exp. Neurol. 2016, 284, 38–49. [Google Scholar] [CrossRef] [PubMed]
  40. Osellame, L.D.; Rahim, A.A.; Hargreaves, I.P.; Gegg, M.E.; Richard-Londt, A.; Brandner, S.; Waddington, S.N.; Schapira, A.H.V.; Duchen, M.R. Mitochondria and Quality Control Defects in a Mouse Model of Gaucher Disease—Links to Parkinson’s Disease. Cell Metab. 2013, 17, 941–953. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Jolly, R.D.; Brown, S.; Das, A.M.; Walkley, S.U. Mitochondrial dysfunction in the neuronal ceroid-lipofuscinoses (Batten disease). Neurochem. Int. 2002, 40, 565–571. [Google Scholar] [CrossRef]
  42. Plotegher, N.; Duchen, M.R. Mitochondrial Dysfunction and Neurodegeneration in Lysosomal Storage Disorders. Trends Mol. Med. 2017, 23, 116–134. [Google Scholar] [CrossRef] [PubMed]
  43. Settembre, C.; Fraldi, A.; Jahreiss, L.; Spampanato, C.; Venturi, C.; Medina, D.; de Pablo, R.; Tacchetti, C.; Rubinsztein, D.C.; Ballabio, A. A block of autophagy in lysosomal storage disorders. Hum. Mol. Genet. 2008, 17, 119–129. [Google Scholar] [CrossRef] [PubMed]
  44. Takamura, A.; Higaki, K.; Kajimaki, K.; Otsuka, S.; Ninomiya, H.; Matsuda, J.; Ohno, K.; Suzuki, Y.; Nanba, E. Enhanced autophagy and mitochondrial aberrations in murine G(M1)-gangliosidosis. Biochem. Biophys. Res. Commun. 2008, 367, 616–622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Martins, C.; Hůlková, H.; Dridi, L.; Dormoy-Raclet, V.; Grigoryeva, L.; Choi, Y.; Langford-Smith, A.; Wilkinson, F.L.; Ohmi, K.; DiCristo, G.; et al. Neuroinflammation, mitochondrial defects and neurodegeneration in mucopolysaccharidosis III type C mouse model. Brain 2015, 138, 336–355. [Google Scholar] [CrossRef] [Green Version]
  46. Selak, M.A.; de Chadarevian, J.P.; Melvin, J.J.; Grover, W.D.; Salganicoff, L.; Kaye, E.M. Mitochondrial activity in Pompe’s disease. Pediatr. Neurol. 2000, 23, 54–57. [Google Scholar] [CrossRef]
  47. Otomo, T.; Higaki, K.; Nanba, E.; Ozono, K.; Sakai, N. Inhibition of autophagosome formation restores mitochondrial function in mucolipidosis II and III skin fibroblasts. Mol. Genet. Metab. 2009, 98, 393–399. [Google Scholar] [CrossRef]
  48. Jennings, J.J., Jr.; Zhu, J.H.; Rbaibi, Y.; Luo, X.; Chu, C.T.; Kiselyov, K. Mitochondrial aberrations in mucolipidosis Type IV. J. Biol. Chem. 2006, 281, 39041–39050. [Google Scholar] [CrossRef] [Green Version]
  49. Sano, R.; Annunziata, I.; Patterson, A.; Moshiach, S.; Gomero, E.; Opferman, J.; Forte, M.; d’Azzo, A. GM1-ganglioside accumulation at the mitochondria-associated ER membranes links ER stress to Ca2+-dependent mitochondrial apoptosis. Mol. Cell 2009, 36, 500–511. [Google Scholar] [CrossRef] [Green Version]
  50. Cao, Y.; Espinola, J.A.; Fossale, E.; Massey, A.C.; Cuervo, A.M.; MacDonald, M.E.; Cotman, S.L. Autophagy is disrupted in a knock-in mouse model of juvenile neuronal ceroid lipofuscinosis. J. Biol. Chem. 2006, 281, 20483–20493. [Google Scholar] [CrossRef] [Green Version]
  51. Yambire, K.F.; Fernandez-Mosquera, L.; Steinfeld, R.; Mühle, C.; Ikonen, E.; Milosevic, I.; Raimundo, N. Mitochondrial biogenesis is transcriptionally repressed in lysosomal lipid storage diseases. eLife 2019, 8. [Google Scholar] [CrossRef]
  52. De Pablo-Latorre, R.; Saide, A.; Polishhuck, E.V.; Nusco, E.; Fraldi, A.; Ballabio, A. Impaired parkin-mediated mitochondrial targeting to autophagosomes differentially contributes to tissue pathology in lysosomal storage diseases. Hum. Mol. Genet. 2012, 21, 1770–1781. [Google Scholar] [CrossRef]
  53. Deus, C.M.; Yambire, K.F.; Oliveira, P.J.; Raimundo, N. Mitochondria–Lysosome Crosstalk: From Physiology to Neurodegeneration. Trends Mol. Med. 2020, 26, 71–88. [Google Scholar] [CrossRef] [Green Version]
  54. Burgoyne, T.; Patel, S.; Eden, E.R. Calcium signaling at ER membrane contact sites. Biochim. Biophys. Acta (BBA) Mol. Cell Res. 2015, 1853, 2012–2017. [Google Scholar] [CrossRef] [Green Version]
  55. Soto-Heredero, G.; Baixauli, F.; Mittelbrunn, M. Interorganelle Communication between Mitochondria and the Endolysosomal System. Front. Cell Dev. Biol. 2017, 5, 95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Raffaello, A.; Mammucari, C.; Gherardi, G.; Rizzuto, R. Calcium at the Center of Cell Signaling: Interplay between Endoplasmic Reticulum, Mitochondria, and Lysosomes. Trends Biochem. Sci. 2016, 41, 1035–1049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Bertero, E.; Maack, C. Calcium Signaling and Reactive Oxygen Species in Mitochondria. Circ. Res. 2018, 122, 1460–1478. [Google Scholar] [CrossRef]
  58. Chang, K.; Niescier, R.F.; Min, K. Mitochondrial matrix Ca2+ as an intrinsic signal regulating mitochondrial motility in axons. Proc. Natl. Acad. Sci. USA 2011, 108, 15456–15461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Baixauli, F.; Acín-Pérez, R.; Villarroya-Beltrí, C.; Mazzeo, C.; Nuñez-Andrade, N.; Gabandé-Rodriguez, E.; Dolores Ledesma, M.; Blázquez, A.; Martin, M.A.; Falcón-Pérez, J.M.; et al. Mitochondrial respiration controls lysosomal function during inflammatory T cell responses. Cell Metab. 2015, 22, 485–498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Ekstrand, M.I.; Falkenberg, M.; Rantanen, A.; Park, C.B.; Gaspari, M.; Hultenby, K.; Rustin, P.; Gustafsson, C.M.; Larsson, N.G. Mitochondrial transcription factor A regulates mtDNA copy number in mammals. Hum. Mol. Genet. 2004, 13, 935–944. [Google Scholar] [CrossRef]
  61. Larsson, N.-G.; Wang, J.; Wilhelmsson, H.; Oldfors, A.; Rustin, P.; Lewandoski, M.; Barsh, G.S.; Clayton, D.A. Mitochondrial transcription factor A is necessary for mtDNA maintance and embryogenesis in mice. Nat. Genet. 1998, 18, 231–236. [Google Scholar] [CrossRef]
  62. Mindell, J.A. Lysosomal Acidification Mechanisms. Annu. Rev. Physiol. 2012, 74, 69–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Trivedi, P.C.; Bartlett, J.J.; Pulinilkunnil, T. Lysosomal Biology and Function: Modern View of Cellular Debris Bin. Cells 2020, 9, 1131. [Google Scholar] [CrossRef] [PubMed]
  64. Montero, R.; Yubero, D.; Salgado, M.C.; González, M.J.; Campistol, J.; O’Callaghan, M.d.M.; Pineda, M.; Delgadillo, V.; Maynou, J.; Fernandez, G.; et al. Plasma coenzyme Q10 status is impaired in selected genetic conditions. Sci. Rep. 2019, 9, 793. [Google Scholar] [CrossRef] [PubMed]
  65. Banerjee, S.; Kane, P.M. Regulation of V-ATPase Activity and Organelle pH by Phosphatidylinositol Phosphate Lipids. Front. Cell Dev. Biol. 2020, 8. [Google Scholar] [CrossRef]
  66. Demers-Lamarche, J.; Guillebaud, G.; Tlili, M.; Todkar, K.; Bélanger, N.; Grondin, M.; Nguyen, A.P.; Michel, J.; Germain, M. Loss of Mitochondrial Function Impairs Lysosomes. J. Biol. Chem. 2016, 291, 10263–10276. [Google Scholar] [CrossRef] [Green Version]
  67. Cerutti, R.; Pirinen, E.; Lamperti, C.; Marchet, S.; Sauve, A.A.; Li, W.; Leoni, V.; Schon, E.A.; Dantzer, F.; Auwerx, J.; et al. NAD+-Dependent Activation of Sirt1 Corrects the Phenotype in a Mouse Model of Mitochondrial Disease. Cell Metab. 2014, 19, 1042–1049. [Google Scholar] [CrossRef] [Green Version]
  68. Gomes, A.P.; Price, N.L.; Ling, A.J.Y.; Moslehi, J.J.; Montgomery, M.K.; Rajman, L.; White, J.P.; Teodoro, J.S.; Wrann, C.D.; Hubbard, B.P.; et al. Declining NAD+ Induces a Pseudohypoxic State Disrupting Nuclear-Mitochondrial Communication during Aging. Cell 2013, 155, 1624–1638. [Google Scholar] [CrossRef] [Green Version]
  69. Karamanlidis, G.; Lee, C.F.; Garcia-Menendez, L.; Kolwicz, S.C.; Suthammarak, W.; Gong, G.; Sedensky, M.M.; Morgan, P.G.; Wang, W.; Tian, R. Mitochondrial Complex I Deficiency Increases Protein Acetylation and Accelerates Heart Failure. Cell Metab. 2013, 18, 239–250. [Google Scholar] [CrossRef] [Green Version]
  70. De la Mata, M.; Cotán, D.; Oropesa-Ávila, M.; Garrido-Maraver, J.; Cordero, M.D.; Villanueva Paz, M.; Delgado Pavón, A.; Alcocer-Gómez, E.; de Lavera, I.; Ybot-González, P.; et al. Pharmacological Chaperones and Coenzyme Q10 Treatment Improves Mutant β-Glucocerebrosidase Activity and Mitochondrial Function in Neuronopathic Forms of Gaucher Disease. Sci. Rep. 2015, 5, 10903. [Google Scholar] [CrossRef] [Green Version]
  71. Heaton, R.A.; Heales, S.; Rahman, K.; Sexton, D.W.; Hargreaves, I. The Effect of Cellular Coenzyme Q10 Deficiency on Lysosomal Acidification. J. Clin. Med. 2020, 9, 1923. [Google Scholar] [CrossRef]
  72. Zhao, R.Z.; Jiang, S.; Zhang, L.; Yu, Z.B. Mitochondrial electron transport chain, ROS generation and uncoupling (Review). Int. J. Mol. Med. 2019, 44, 3–15. [Google Scholar] [CrossRef] [Green Version]
  73. Lotfi, P.; Tse, D.Y.; Di Ronza, A.; Seymour, M.L.; Martano, G.; Cooper, J.D.; Pereira, F.A.; Passafaro, M.; Wu, S.M.; Sardiello, M. Trehalose reduces retinal degeneration, neuroinflammation and storage burden caused by a lysosomal hydrolase deficiency. Autophagy 2018, 14, 1419–1434. [Google Scholar] [CrossRef] [Green Version]
  74. Pierzynowska, K.; Gaffke, L.; Podlacha, M.; Brokowska, J.; Węgrzyn, G. Mucopolysaccharidosis and Autophagy: Controversies on the Contribution of the Process to the Pathogenesis and Possible Therapeutic Applications. Neuromol. Med. 2020, 22, 25–30. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Platt, F.M.; Boland, B.; van der Spoel, A.C. The cell biology of disease: Lysosomal storage disorders: The cellular impact of lysosomal dysfunction. J. Cell Biol. 2012, 199, 723–734. [Google Scholar] [CrossRef] [Green Version]
  76. McGuire, C.; Cotter, K.; Stransky, L.; Forgac, M. Regulation of V-ATPase assembly and function of V-ATPases in tumor cell invasiveness. Biochim. Biophys. Acta 2016, 1857, 1213–1218. [Google Scholar] [CrossRef] [PubMed]
  77. Kang, H.T.; Park, J.T.; Choi, K.; Kim, Y.; Choi, H.J.C.; Jung, C.W.; Lee, Y.S.; Park, S.C. Chemical screening identifies ATM as a target for alleviating senescence. Nat. Chem. Biol. 2017, 13, 616–623. [Google Scholar] [CrossRef] [PubMed]
  78. Stepien, K.M.; Roncaroli, F.; Turton, N.; Hendriksz, C.J.; Roberts, M.; Heaton, R.A.; Hargreaves, I. Mechanisms of Mitochondrial Dysfunction in Lysosomal Storage Disorders: A Review. J. Clin. Med. 2020, 9, 2596. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the lysosome and autophagy pathway in normal cells (A) and lysosomal storage disease (LSD) cells (B). (A) Lysosomes fuse with autophagosomes to form autolysosomes. Lysosomes also fuse with endosomes to form highly dynamic membrane structures called endolysosomes. Functional lysosomes require an acidic environment in the lysosomal lumen (pH 4.5 to 5.0) maintained by the V-ATPase proton pump and an adequate level of calcium ions maintained by the TRPML1 transporter. TRPML1: mammalian mucolipin TRP channel subfamily, V-ATPase: a vacuolar ATPase. (B) LSD is a metabolic disorder caused by mutations in genes that encode lysosomal enzymes, consequently leading to accumulation of various substrates. Lysosomes loaded with undigested substances lead to defective lysosomal fusion with autophagosomes or endosomes. Furthermore, lysosomes of LSD exhibit alterations in lysosomal pH and lysosomal Ca2+ homeostasis.
Figure 1. Schematic representation of the lysosome and autophagy pathway in normal cells (A) and lysosomal storage disease (LSD) cells (B). (A) Lysosomes fuse with autophagosomes to form autolysosomes. Lysosomes also fuse with endosomes to form highly dynamic membrane structures called endolysosomes. Functional lysosomes require an acidic environment in the lysosomal lumen (pH 4.5 to 5.0) maintained by the V-ATPase proton pump and an adequate level of calcium ions maintained by the TRPML1 transporter. TRPML1: mammalian mucolipin TRP channel subfamily, V-ATPase: a vacuolar ATPase. (B) LSD is a metabolic disorder caused by mutations in genes that encode lysosomal enzymes, consequently leading to accumulation of various substrates. Lysosomes loaded with undigested substances lead to defective lysosomal fusion with autophagosomes or endosomes. Furthermore, lysosomes of LSD exhibit alterations in lysosomal pH and lysosomal Ca2+ homeostasis.
Cells 10 00420 g001
Figure 2. Schematic representation of the basic mechanisms of mitochondrial homeostasis in normal cells (A) and LSD-related mitochondrial dysfunction in LSD cells (B). (A) Cytoplasmic Ca2+ modulates GTPase activity, which regulates the balance of mitochondrial fusion/fission events. ER: endoplasmic reticulum. (B) Dysregulation of lysosomal Ca2+ homeostasis in LSD interferes with the balance of mitochondrial fission and fusion cycles. Lysosomal deficiency deteriorates Ca2+ buffering.
Figure 2. Schematic representation of the basic mechanisms of mitochondrial homeostasis in normal cells (A) and LSD-related mitochondrial dysfunction in LSD cells (B). (A) Cytoplasmic Ca2+ modulates GTPase activity, which regulates the balance of mitochondrial fusion/fission events. ER: endoplasmic reticulum. (B) Dysregulation of lysosomal Ca2+ homeostasis in LSD interferes with the balance of mitochondrial fission and fusion cycles. Lysosomal deficiency deteriorates Ca2+ buffering.
Cells 10 00420 g002
Figure 3. Importance of the mitochondrial–lysosomal axis in LSD. (A) Schematic representation of the basic mechanisms describing the effect of lysosomal function on mitochondria, in a contact-independent or dependent manner. MCU: mitochondrial calcium uniporter. (B) Schematic representation of the underlying mechanisms explaining the deletion effect of the mitochondrial transcription factor Tfam on lysosomal function. Tfam: mitochondrial transcription factor A. (C) Schematic representation of the underlying mechanisms explaining the effect of low ATP production due to mitochondrial dysfunction on lysosomal acidification.
Figure 3. Importance of the mitochondrial–lysosomal axis in LSD. (A) Schematic representation of the basic mechanisms describing the effect of lysosomal function on mitochondria, in a contact-independent or dependent manner. MCU: mitochondrial calcium uniporter. (B) Schematic representation of the underlying mechanisms explaining the deletion effect of the mitochondrial transcription factor Tfam on lysosomal function. Tfam: mitochondrial transcription factor A. (C) Schematic representation of the underlying mechanisms explaining the effect of low ATP production due to mitochondrial dysfunction on lysosomal acidification.
Cells 10 00420 g003
Figure 4. Strategies to activate the mitochondrial–lysosomal axis in LSD. (A) The strategy of activating the mitochondrial–lysosomal axis through coenzyme Q10 (CoQ) supplementation improves mitochondrial function to restore lysosome function. Green and pink CoQ indicate low and high CoQ content, respectively. ROS: reactive oxygen species. (B) The strategy of activating the mitochondrial–lysosomal axis through autophagy activation with trehalose improves pathological symptoms associated with LSD. (C) The strategy of activating the mitochondrial–lysosomal axis through lysosomal re-acidification restores the activity of lysosomal enzymes and increases autophagic flux with restoration of mitochondrial function.
Figure 4. Strategies to activate the mitochondrial–lysosomal axis in LSD. (A) The strategy of activating the mitochondrial–lysosomal axis through coenzyme Q10 (CoQ) supplementation improves mitochondrial function to restore lysosome function. Green and pink CoQ indicate low and high CoQ content, respectively. ROS: reactive oxygen species. (B) The strategy of activating the mitochondrial–lysosomal axis through autophagy activation with trehalose improves pathological symptoms associated with LSD. (C) The strategy of activating the mitochondrial–lysosomal axis through lysosomal re-acidification restores the activity of lysosomal enzymes and increases autophagic flux with restoration of mitochondrial function.
Cells 10 00420 g004
Figure 5. Synoptic representation of the mitochondrial–lysosomal axis contributing to the control of LSD. The activation of the mitochondrial–lysosomal axis through mitochondrial functional recovery or lysosomal functional recovery represents a promising and potent strategy to treat LSD.
Figure 5. Synoptic representation of the mitochondrial–lysosomal axis contributing to the control of LSD. The activation of the mitochondrial–lysosomal axis through mitochondrial functional recovery or lysosomal functional recovery represents a promising and potent strategy to treat LSD.
Cells 10 00420 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kuk, M.U.; Lee, Y.H.; Kim, J.W.; Hwang, S.Y.; Park, J.T.; Park, S.C. Potential Treatment of Lysosomal Storage Disease through Modulation of the Mitochondrial—Lysosomal Axis. Cells 2021, 10, 420. https://doi.org/10.3390/cells10020420

AMA Style

Kuk MU, Lee YH, Kim JW, Hwang SY, Park JT, Park SC. Potential Treatment of Lysosomal Storage Disease through Modulation of the Mitochondrial—Lysosomal Axis. Cells. 2021; 10(2):420. https://doi.org/10.3390/cells10020420

Chicago/Turabian Style

Kuk, Myeong Uk, Yun Haeng Lee, Jae Won Kim, Su Young Hwang, Joon Tae Park, and Sang Chul Park. 2021. "Potential Treatment of Lysosomal Storage Disease through Modulation of the Mitochondrial—Lysosomal Axis" Cells 10, no. 2: 420. https://doi.org/10.3390/cells10020420

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

Article Metrics

Back to TopTop