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 -- 3996 2023-07-01 15:52:54 |
2 format change Meta information modification 3996 2023-07-03 05:03:05 |

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.
Misra, S.K.; Rosenholm, J.M.; Pathak, K. Nanosystems for Mitochondrial Drug Delivery with Metallic Nanoparticles. Encyclopedia. Available online: https://encyclopedia.pub/entry/46305 (accessed on 27 July 2024).
Misra SK, Rosenholm JM, Pathak K. Nanosystems for Mitochondrial Drug Delivery with Metallic Nanoparticles. Encyclopedia. Available at: https://encyclopedia.pub/entry/46305. Accessed July 27, 2024.
Misra, Shashi Kiran, Jessica M. Rosenholm, Kamla Pathak. "Nanosystems for Mitochondrial Drug Delivery with Metallic Nanoparticles" Encyclopedia, https://encyclopedia.pub/entry/46305 (accessed July 27, 2024).
Misra, S.K., Rosenholm, J.M., & Pathak, K. (2023, July 01). Nanosystems for Mitochondrial Drug Delivery with Metallic Nanoparticles. In Encyclopedia. https://encyclopedia.pub/entry/46305
Misra, Shashi Kiran, et al. "Nanosystems for Mitochondrial Drug Delivery with Metallic Nanoparticles." Encyclopedia. Web. 01 July, 2023.
Nanosystems for Mitochondrial Drug Delivery with Metallic Nanoparticles
Edit

The application of metallic nanoparticles as a novel therapeutic tool has significant potential to facilitate the treatment and diagnosis of mitochondria-based disorders. Subcellular mitochondria have been trialed to cure pathologies that depend on their dysfunction. Nanoparticles made from metals and their oxides (including gold, iron, silver, platinum, zinc oxide, and titanium dioxide) have unique modi operandi that can competently rectify mitochondrial disorders.

metallic nanoparticles mitochondrial dysfunction antioxidants reactive oxygen species calcium homeostasis biocompatibility

1. Introduction

Mitochondria, one of the vital cell organelles, is notable as being oval or bean shaped, as well as for containing a dynamic branched system that consistently divides and fuses in accordance with the regulation of fission and fusion theory [1]. Mitochondria are the energy generator of eukaryotic cells, in addition to producing the adenosine triphosphate (ATP) that performs diverse functions, including the biosynthesis and degradation of protein molecules, cell division and respiration, as well as controlling membrane potential. They are adorned with countless cristae in the inner membrane of mitochondria, which are specifically involved in the generation of ATP on a massive scale. Besides the generation of ATP, mitochondria also perform numerous other tasks, such as the biogenesis of amino acids, ROS signaling, calcium ion homeostasis, apoptosis, stem cell monitoring, and the control of innate immunity [2]. Further, the bidirectional drive between the nucleus and ATP generator is tightly regulated via the fission (inner membrane) and fusion (outer membrane) of mitochondria. Diverse physical platforms (including signaling pathways, protein–protein interaction, and the regulation of both calcium homeostasis and released reactive oxygen species) occur in the powerhouse of the cell. A number of metabolic processes, such as glycolysis, the Krebs cycle, oxidative phosphorylation, and acetyl CoA oxidation, are executed in the mitochondria, both for the purpose of ATP generation, as well as for dumping released electrons from the various pathways.
The mitochondria has an extremely flexible ultrastructure, intended for the regulation of the bioenergetics flux of the cell. The literature maintains that the sequence of mitochondrial genome (MtDNA) can stabilize maternally inherited diseases. Additionally, the proteomics of mitochondria has described approximately 1000 proteins inside the nuclear genome, of which a mere 13 proteins are coded by the mitochondrial DNA. Hence, quality control of mitochondria is critically important to avoid genome defects, as well as to maintain cell metabolic homeostasis [3]. The production, regulation, and consumption of organic molecules are highly required for supporting cell growth and proliferation. Mitochondrial oxygen utilization is key process that generates ATP from these organic compounds and other valuable intermediates. Besides ATP generation, reactive oxygen species (ROS) are also produced during the process of oxidative phosphorylation. These ROS mediate the signaling pathways associated with several basal and adaptive responses controlling both cell and organism hemostasis [4].

2. Mitochondrial Dysfunction and Diseases

Human cells comprise more than a thousand replicas of mitochondrial DNA (MtDNA). At the time of birth, these are homoplasmy that may be altered due to pathogens or mutation, and can consequently develop wild type MtDNA, or heteroplasm. The threshold of pathogenic MtDNA varies from person to person, and from organ to organ in the same individual. Abnormal mitochondria dynamics (i.e., fission and fusion), reduced oxidative phosphorylation capacity, and a transformed electron transport chain can each elicit ATP scarcity, alter ROS with poor membrane potential, and cause specific stress signaling responses. The resultant poor mitochondrial protein transportation influences enzyme assembly, which initiates several consequences, such as oxidative damage, proteotoxic stress, mutation, and the exhaustion of MtDNA. The altered morphology and functions of mitochondria can interrupt their nutrient requirements, substrate availability, and genetic signals, causing inherited MtDNA mutations which thus weaken its defense system. Figure 1 outlines the prominent features of the outer, inner, associated plasma membrane and cytosolic barriers.
Figure 1. Interconnected distinguishing features and processes occurring in the mitochondria and associated sites.
Clinical studies have revealed that progressive aging is implicated with mitochondrial impairment, and that the release and accumulation of ROS plays a crucial role in pathogenesis of age-related issues, including neurodegenerative disorders and diabetes. The imbalanced level of generation and oxidation of ROS disturbs the respiratory chain function of mitochondria, altering not only outer/inner membrane permeability, but also calcium homeostasis and high heteroplasmic MtDNA in the sufferers [5]. Mitochondrial abnormality (either by MtDNA or nDNA) may occur at any age, and can affect single or multiple organs, e.g., the eye and ear. Both Leber hereditary optic neuropathy and nonsyndromic hearing loss (or deafness) have been reported as being due to abnormal mitochondrial functions. The literature reports noticeable clinical features, such as pigmentary retinopathy, cardiomyopathy, diabetes mellitus, and disorders in central nervous system (i.e., ataxia, dementia, seizures, fluctuating encephalopathy, and migraine), as being associated with the abnormal functioning of mitochondria. Figure 2 illustrates primary and secondary mitochondrial disorders and their operational issues.
Figure 2. Mitochondrial disorders (primary and secondary) and their operational issues.
Recently, Friedreich’s ataxia (a neurodegenerative defect) has been identified due to mitochondrial imbalance of ROS and reactive nitrogen species (RNS). One of the mitochondrial proteins, named frataxin, is encoded and arranged by the FXN gene, which is tremendously expressed in tissues with high metabolic rates. Frataxin is vitally involved in accumulation of the iron–sulfur clusters required for energy production. The latest evidence proposes that this protein deactivates the glutathione peroxidase enzyme, elevates thiol levels, and detoxifies the ROS associated with Friedreich’s ataxia. Hence, targeting this protein may be a possible approach for a drug delivery system [6]. Genetic disorders such as Wilson’s disease, hereditary spastic paraplegia, and ethylmalonic aciduria are caused by the mutation of ATP7b, SPG7, and ETHE1 genes, respectively [7]. Table 1 compiles different diseases caused by damaged mitochondria, as well as the primary methods of their management.
Table 1. Mitochondrial abnormalities and their management in different human diseases.

3. Therapeutic Approaches

Approaches such as ‘one size fits all’ and ‘precision medicine’ are often employed for the mitigation of primary mitochondrial disorders. The former manages symptomatic interventions (based on diet, exercise, and pharmacological therapeutics) that initiate mitochondrial biogenesis, provoke nitric acid synthetase, amplify ATP synthesis, control mitochondrial autophagy, and stimulate fission/fusion processes. On the other hand, the precision medicine approach focuses on specialized therapies, i.e., nucleotide supplementation, swapping of damaged MtDNA, gene replacement remedies, the elimination of noxious metabolites, and organ transplantation, if required. Table 2 compiles several therapeutic agents for the regulation of the myriad pathways involved in impaired mitochondria.
Table 2. Therapeutic agents explored for the management of impaired mitochondria.
These approaches may represent a common tactic for the management of primary mitochondrial disorders and can improve an individual’s quality of life, but they are unable to eradicate the cause of the mitochondrial disorder completely.

4. Biological Barrier and Toxicity

There are myriad barriers to designing mitochondria-targeted drug delivery, such as biological barriers and drug toxicity. After reaching to the target cell, intracellular diffusion across the outer and inner mitochondrial membrane is another challenge. Further, mitochondrial complex (I–IV) inhibition within ETC from drugs such as haloperidol and thiothixene have shown critical toxicities [26]. Resultantly, neurological disorder such as Parkinson’s disease arise. Nanotechnology-based approaches are promising tools for effectively targeting mitochondria, owing to their narrow size and their efficiency in transporting bioactive agents. Both the mitochondrial membranes experience variable permeability, i.e., the outer membrane allows the permeation of a large array of small molecules, whereas the inner membrane is quite selective for transportation. The reason behind selective permeability through the inner mitochondrial membrane is the presence of specialized channel proteins there [27]. Further, the transition pore of the outer mitochondrial membrane is wider than the inner mitochondrial membrane, which facilitates the faster traversal of therapeutics. The highly folded inner mitochondrial membrane is hard and possesses narrow transition slits to separate out the mitochondrial matrix within the inter membrane space. This characteristic morphology of mitochondria also poses a barrier for the passage of drug molecules across matrix. The aforesaid issues create hurdles for targeting mitochondria-based disorders. Several aspects, including high negative membrane potential (−160 to −180 mV), unambiguous protein import mechanism, variable lipophilicity, and idiosyncratic phospholipid composition of inner membrane, are a few key factors considered while designing a mitochondrion-targeted drug delivery system [28]. Permeation of the outer membrane of mitochondria is concentration dependent, following passive diffusion. A negative surface charge and mitochondrial membrane potentially drive the production of ATP, in addition to exploiting for mitochondrial targeting through cationic ions and drug molecules via electrostatic interaction. Cardiolipin, another exclusive phospholipid-based component, is confined and produced in the inner mitochondria region. It plays a central role in processing several reactions and processes required for mitochondrial dynamics and respiratory chain complexes [29]. Moreover, electrochemical composition and membrane potential are remarkably distinct, compared to cytoplasm. A mitochondrion comprises its own genome in form of a circular DNA containing 37 genes and 16,500 base pairs, essential for protein synthesis. The synthesized 13 proteins are basic components of the mitochondrial electron transport chain, as well as performing the process of oxidative phosphorylation [30]. Mitohormesis is related to the damaged adaptive response of mitochondria, and was very first discussed by Ristow et al. in 2010. This process controls mitochondrial homoeostasis and assists organismal senescence. The literature supports that increased age favors defects in mitochondria (mitophagy), replacing new mitochondria via the process of fission and biogenesis [31]. Therefore, the mitochondria’s basic functions deteriorate with aging, and can bring about several age-related disorders. In this context, drug-induced cell and mitochondrial toxicities are extensively demarcated in the geriatric population into myriad categories of pharmacological agents [32]. For example, troglitazone (an antidiabetic drug) exhibited mitochondrial toxicity and was withdrawn from the market after liver dysfunction was reported. The drug potently disrupted mitochondrial-oriented respiration by inhibiting complexes II, III, IV, and V [33]. Another, nefazodone (a serotonin antagonist) has been withdrawn from the market in United States in the year 2004 due to profoundly inhibiting mitochondrial respiratory chain in isolated rat liver cells [34]. Biguanides (buformin, metformin, and phenformin) trigger lactic acidosis, which is directly associated with mitochondrial damage due to a surge in lactate production and respiratory inhibition [35]. An over-the-counter drug, aspirin is the most prescribed medicine in elderly patients for overcoming of pain and has been reported to inhibit respiratory processes, as well as opening MPT pores and disturbing of glutathione status in mitochondria [36]. Another cholesterol lowering drug, simvastatin, obstructs the ETC complexes I, IV, and V, augments calcium release, reduces mitochondrial membrane potential, and thus decreases ATP levels [37]. Further, both the in vivo administration of impoverished pharmacokinetic properties and the erratic biodistribution of drugs create challenges for the design of delivery systems. In many instances, therapeutic agents lost their potency with the addition of selective coding for the targeting of mitochondria.

5. Nanoengineered Mitochondria Targeted Approaches

Nanotechnology is able to advance the pharmacokinetic and biodistribution profiles of various drug molecules without altering their pristine molecular form. The nanoformulation of drugs can modify the inherent physicochemical attributes of drugs, such as solubility, lipophilicity, half-life, and minimization of immunogenicity. Successful delivery of bioactive substances with the use of nanoparticles can be realized by customizing particle size, lipophilicity, surface charge, and the addition of specific targeting moieties. Being biocompatible and biodegradable, nanoparticles have proven potential for site-specific delivery [38], in addition to the intracellular targeting of subcellular compartments. It is evident that the architecture of mitochondria is highly distinct from other subcellular parts in eukaryotic cells. The occurrence of proton pumps in the inner mitochondrial membrane creates positive charges on the inner folded surface, whereas a negative charge is found in the mitochondrial matrix. This charge gradient develops noticeable transmembrane potential across both outer and inner mitochondrial membranes [39]. Molecules such as pyruvates and lipophilic cations (i.e., triphenylphosphonium, rhodamine123, and tetrachlorotetraethylimidacarbocyanine) have been reported on for targeting mitochondria. The cationic lipoidal content of these molecules enables their permeation into the mitochondrial matrix, thus effectively targeting the region of interest. Positively charged nanoparticles induce an electrostatic interaction with the anionic phospholipids of mitochondrial membranes, and are then internalized. Inside the membrane, the nanoparticles rupture and release their drug with the mitochondrial matrix [40].
Recently, the cationic liposome-like vesicles known as ‘DQAsomes’ have been highly researched as mitochondria-targeted carriers, being explored to deliver cytotoxic therapeutics or DNA into the highly negative environment of cancer cells [41], as these cells experience hyperpolarized membrane potential (−220 mV) compared to the mitochondrial membrane potential of normal cells (−140 mV). Bolalipids containing DQAsomes are vesicle-like structures that are tuned for the delivery of DNA, genes, and peptides inside the mitochondrial matrix. The entrapped component is selectively transported via adorned protein pump machinery on the surface of mitochondria. DQAplexes are hybrid assemblies, comprising membrane liposomes with plasmid DNA for delivery of chemotherapeutics (specifically, into the mitochondria). Vaidya et al. conjugated folic acid on the surface of paclitaxel-entrapped DQAsomes to target overexpressive folate receptors on HeLa (tumor) cells. Outcomes obtained from confocal laser scanning microscopy revealed the enhanced antitumor activity of the designed functionalized DQAsomes, compared to bare DQAsomes [42]. However, the modi operandi for restricting the precise mitochondrial targeting of this system were unclear. Other nanoengineered dendrimers with high generation number are abundantly promoted to deliver cytotoxic agents to the mitochondria. Cations (such as triphenylphosphonium, polyamidoamine, and rhodamine) containing dendrimers encompass high net-positive charges and had the ability to carry genetic materials and chemotherapeutic agents at the cell organelle [43]. Table 3 summarizes the various nanodrug delivery systems utilized for the management of damaged mitochondria-based diseases.
Table 3. Nanoengineered systems for mitigation of impaired mitochondrial diseases.
Metal-based nanoparticles of sizes ranging between 5 and 260 nm were suitable for entry into the mitochondrial matrix. Wang et al. synthesized gold nanorods, in addition to discussing the comparative cellular internalization and intracellular trafficking through the mitochondria, lysosomes, cytoplasm, and endosomes of both normal and cancerous cells. A seed-mediated growth method was utilized to design cetyltrimethylammonium–ammonium bromide wrapped gold nanorods. The average particle size and zeta potential of developed nanorods were 18 ± 0.5 nm and +29.3 ± 0.7 mV, respectively. The prolonged retention and cationic surface charge of the nanorods reduced the mitochondrial membrane potential and generated reactive oxygen species in the mitochondria of lung cancerous cells (A549 cells), which caused apoptosis or cell death [51]. Figure 3 outlines the principal mechanism of metallic nanoparticles for targeting mitochondria.
Figure 3. Schematic illustration on the mechanisms of metallic nanoparticles for targeting different pathway signals of mitochondria. PI3K: Phosphatidylinositol 3 Kinase; mTOR: Mammalian Target of Rapamycin; AMPK: Adenosine Monophosphate Activated Protein Kinase; Erk: Extracellular Signal-Regulated Kinase; Akt: Protein Kinase B; MAPK: Mitogen Activated Protein Kinase; ULK1: 112-kDa protein; BNIP3/NIX: Bcl-2 interacting protein 3; PINK1/Parkin-PTEN-induced kinase 1/autoinhibited E3-Ub ligase; SQSTM/P62: Sequestosome 1/p62 scaffold protein; NBR1: Ubiquitous Scaffold Protein; LC3-II: Ubiquitin-Binding Domain (Phosphatidylethanolamine conjugate).
However, the toxicity of metallic nanoparticles is reliant on biophysical characteristics, such as size, surface area, charge, and aggregation. These features affect biodistribution and internalization within organ systems, in addition to changing molecular interactions with receptors or macromolecules [52]. Published reports have suggested that there is a direct correlation between the particle size of nanoparticles and their efficiency to generate ROS in vital organs. For instance, silver nanoparticles with 10 nm size displayed superior biodistribution and caused more fatal effects in the liver and spleen cells, compared to the size range 40–100 nm [53]. As with the nanoparticle size, shape also affect biodistribution and clearance from the body. Long fibrous metallic nanoparticles exhibited more serious effects and were difficult to remove out from organ systems [54]. Similarly charged nanoparticles show greater accumulation in the target site compared to the noncharged ones. Peak et al. investigated the effects of particle size and surface charge of zinc oxide nanoparticles on pharmacokinetic parameters, including biodistribution and clearance upon administration of a single oral dose to rats. It was evident that negatively charged zinc oxide nanoparticles were extensively absorbed by the systemic circulation, compared to those that were positively charged. Zinc oxide particles of 20 nm were swiftly eliminated through the biliary and fecal routes [55].

References

  1. Mishra, P.; Chan, D.C. Mitochondrial dynamics and inheritance during cell division, development and disease. Nat. Rev. Mol. Cell Biol. 2014, 15, 634–646.
  2. Kühlbrandt, W. Structure and function of mitochondrial membrane protein complexes. BMC Biol. 2015, 13, 89.
  3. Calvo, S.E.; Mootha, V.K. The mitochondrial proteome and human disease. Annu. Rev. Genom. Hum. Genet. 2010, 11, 25–44.
  4. Shadel, G.S.; Horvath, T.L. Mitochondrial ROS signaling in organismal homeostasis. Cell 2015, 163, 560–569.
  5. Elfawy, H.A.; Das, B. Crosstalk between mitochondrial dysfunction, oxidative stress, and age-related neurodegenerative disease: Etiologies and therapeutic strategies. Life Sci. 2019, 218, 165–184.
  6. Calabrese, V.; Lodi, R.; Tonon, C. D’Agata. V.; Sapienza, M.; Scapagnini, G.; Mangiameli, A.; Pennisi, G.; Stella, A.M.; Butterfield, D.A. Oxidative stress, mitochondrial dysfunction and cellular stress response in Friedreich’s ataxia. J. Neurol. Sci. 2005, 233, 145–162.
  7. Hellebrekers, D.M.; Wolfe, R.; Hendrickx, A.T.; de Coo, I.F.; de Die, C.E.; Geraedts, J.P.; Chinnery, P.F.; Smeets, H.J. PGD and heteroplasmic mitochondrial DNA point mutations: A systematic review estimating the chance of healthy offspring. Hum. Reprod. Update 2012, 18, 341–349.
  8. Bisaccia, G.; Ricci, F.; Gallina, S.; Di Baldassarre, A.; Ghinassi, B. Mitochondrial Dysfunction and Heart Disease: Critical Appraisal of an Overlooked Association. Int. J. Mol. Sci. 2021, 22, 614.
  9. Sergi, D.; Naumovski, N.; Heilbronn, L.K.; Abeywardena, M.; O’Callaghan, N.; Lionetti, L.; Luscombe-Marsh, N. Mitochondrial (Dys)function and Insulin Resistance: From Pathophysiological Molecular Mechanisms to the Impact of Diet. Front. Physiol. 2019, 10, 532.
  10. Takemura, K.; Nishi, H.; Inagi, R. Mitochondrial Dysfunction in Kidney Disease and Uremic Sarcopenia. Front. Physiol. 2020, 11, 565023.
  11. Misrani, A.; Tabassum, S.; Yang, L. Mitochondrial Dysfunction and Oxidative Stress in Alzheimer’s Disease. Front. Aging Neurosci. 2021, 13, 617588.
  12. Liu, Y.; Shi, Y. Mitochondria as a target in cancer treatment. MedComm 2020, 1, 129–139.
  13. Kamatani, N.; Kushiyama, A.; Toyo-Oka, L.; Toyo-Oka, T. Treatment of two mitochondrial disease patients with a combination of febuxostat and inosine that enhances cellular ATP. J. Hum. Genet. 2019, 64, 351–353.
  14. Viscomi, C.; Bottani, E.; Civiletto, G.; Cerutti, R.; Moggio, M.; Fagiolari, G.; Schon, E.A.; Lamperti, C.; Zeviani, M. In vivo correction of COX deficiency by activation of the AMPK/PGC-1α axis. Cell. Metab. 2011, 14, 80–90.
  15. Golubitzky, A.; Dan, P.; Weissman, S.; Link, G.; Wikstrom, J.D.; Saada, A. Screening for active small molecules in mitochondrial complex I deficient patient’s fibroblasts, reveals AICAR as the most beneficial compound. PLoS ONE 2011, 6, e26883.
  16. Airhart, S.E.; Shireman, L.M.; Risler, L.J.; Anderson, G.D.; Nagana Gowda, G.A.; Raftery, D.; Tian, R.; Shen, D.D.; O’Brien, K.D. An open-label, non-randomized study of the pharmacokinetics of the nutritional supplement nicotinamide riboside (NR) and its effects on blood NAD+ levels in healthy volunteers. PLoS ONE 2017, 12, e0186459.
  17. Sudo, A.; Sano, H.; Kawamura, N. Determination of the critical time point for efficacy of L-arginine infusion therapy in a case of MELAS with frequent stroke-like episodes. No Hattatsu 2014, 46, 39–43.
  18. Lorenz, C.; Lesimple, P.; Bukowiecki, R.; Zink, A.; Inak, G.; Mlody, B.; Singh, M.; Semtner, M.; Mah, N.; Auré, K.; et al. Human iPSC-Derived Neural Progenitors Are an Effective Drug Discovery Model for Neurological mtDNA Disorders. Cell Stem Cell 2017, 20, 659–674.e9.
  19. Bartsakoulia, M.; Mϋller, J.S.; Gomez-Duran, A.; Yu-Wai-Man, P.; Boczonadi, V.; Horvath, R. Cysteine Supplementation May be Beneficial in a Subgroup of Mitochondrial Translation Deficiencies. J. Neuromuscul. Dis. 2016, 3, 363–379.
  20. Guha, S.; Konkwo, C.; Lavorato, M.; Mathew, N.D.; Peng, M.; Ostrovsky, J.; Kwon, Y.J.; Polyak, E.; Lightfoot, R.; Seiler, C.; et al. Pre-clinical evaluation of cysteamine bitartrate as a therapeutic agent for mitochondrial respiratory chain disease. Hum. Mol. Genet. 2019, 28, 1837–1852.
  21. Blankenberg, F.G.; Kinsman, S.L.; Cohen, B.H.; Goris, M.L.; Spicer, K.M.; Perlman, S.L.; Krane, E.J.; Kheifets, V.; Thoolen, M.; Miller, G.; et al. Brain uptake of Tc99m-HMPAO correlates with clinical response to the novel redox modulating agent EPI-743 in patients with mitochondrial disease. Mol. Genet. Metab. 2012, 107, 690–699.
  22. Kouga, T.; Takagi, M.; Miyauchi, A.; Shimbo, H.; Iai, M.; Yamashita, S.; Murayama, K.; Klein, M.B.; Miller, G.; Goto, T.; et al. Japanese Leigh syndrome case treated with EPI-743. Brain Dev. 2018, 40, 145–149.
  23. Fabbri, A.; Travaglione, S.; Maroccia, Z.; Guidotti, M.; Pierri, C.L.; Primiano, G.; Servidei, S.; Loizzo, S.; Fiorentini, C. The Bacterial Protein CNF1 as a Potential Therapeutic Strategy against Mitochondrial Diseases: A Pilot Study. Int. J. Mol. Sci. 2018, 19, 1825.
  24. Johnson, S.C.; Yanos, M.E.; Kayser, E.B.; Quintana, A.; Sangesland, M.; Castanza, A.; Uhde, L.; Hui, J.; Wall, V.Z.; Gagnidze, A.; et al. mTOR inhibition alleviates mitochondrial disease in a mouse model of Leigh syndrome. Science 2013, 342, 1524–1528.
  25. Sanz, A.; Soikkeli, M.; Portero-Otín, M.; Wilson, A.; Kemppainen, E.; McIlroy, G.; Ellilä, S.; Kemppainen, K.K.; Tuomela, T.; Lakanmaa, M.; et al. Expression of the yeast NADH dehydrogenase Ndi1 in Drosophila confers increased lifespan independently of dietary restriction. Proc. Natl. Acad. Sci. USA 2010, 107, 9105–9110.
  26. Burkhardt, C.; Kelly, J.P.; Lim, Y.H.; Filley, C.M.; Parker, W.D., Jr. Neuroleptic medications inhibit complex I of the electron transport chain. Ann. Neurol. 1993, 33, 512–517.
  27. Pathak, R.K.; Kolishetti, N.; Dhar, S. Targeted nanoparticles in mitochondrial medicine. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2015, 7, 315–329.
  28. Yu, H.; Koilkonda, R.D.; Chou, T.-H.; Porciatti, V.; Ozdemir, S.S.; Chiodo, V.; Boye, S.L.; Boye, S.E.; Hauswirth, W.W.; Lewin, A.S. Gene delivery to mitochondria by targeting modified adenoassociated virus suppresses Leber’s hereditary optic neuropathy in a mouse model. Proc. Natl. Acad. Sci. USA 2012, 109, E1238–E1247.
  29. Kagan, V.E.; Borisenko, G.G.; Tyurina, Y.Y.; Tyurin, V.A.; Jiang, J.; Potapovich, A.I.; Kini, V.; Amoscato, A.A.; Fujii, Y. Oxidative lipidomics of apoptosis: Redox catalytic interactions of cytochrome c with cardiolipin and phosphatidylserine. Free. Radic. Biol. Med. 2004, 37, 1963–1985.
  30. Wallace, D.C.; Fan, W. Energetics, epigenetics, mitochondrial genetics. Mitochondrion 2010, 10, 12–31.
  31. Ristow, M.; Zarse, K. How increased oxidative stress promotes longevity and metabolic health: The concept of mitochondrial hormesis (mitohormesis). Exp. Gerontol. 2010, 45, 410–418.
  32. Will, Y.; Shields, J.E.; Wallace, K.B. Drug-Induced Mitochondrial Toxicity in the Geriatric Population: Challenges and Future Directions. Biology 2019, 8, 32.
  33. Nadanaciva, S.; Dykens, J.A.; Bernal, A.; Capaldi, R.A.; Will, Y. Mitochondrial impairment by PPAR agonists and statins identified via immunocaptured OXPHOS complex activities and respiration. Toxicol. Appl. Pharmacol. 2007, 223, 277–287.
  34. Dykens, J.A.; Jamieson, J.D.; Marroquin, L.D.; Nadanaciva, S.; Xu, J.J.; Dunn, M.C.; Smith, A.R.; Will, Y. In vitro assessment of mitochondrial dysfunction and cytotoxicity of nefazodone, trazodone, and buspirone. Toxicol. Sci. 2008, 103, 335–345.
  35. Dykens, J.A.; Jamieson, J.; Marroquin, L.; Nadanaciva, S.; Billis, P.A.; Will, Y. Biguanide-induced mitochondrial dysfunction yields increased lactate production and cytotoxicity of aerobically poised HepG2 cells and human hepatocytes in vitro. Toxicol. Appl. Pharmacol. 2008, 233, 203–210.
  36. Raza, H.; John, A. Implications of altered glutathione metabolism in aspirin-induced oxidative stress and mitochondrial dysfunction in HepG2 cells. PLoS ONE 2012, 7, e36325.
  37. Bonifacio, A.; Mullen, P.J.; Mityko, I.S.; Navegantes, L.C.; Bouitbir, J.; Krähenbühl, S. Simvastatin induces mitochondrial dysfunction and increased atrogin-1 expression in H9c2 cardiomyocytes and mice in vivo. Arch. Toxicol. 2016, 90, 203–215.
  38. Marrache, S.; Pathak, R.K.; Darley, K.L.; Choi, J.H.; Zaver, D.; Kolishetti, N.; Dhar, S. Nanocarriers for tracking and treating diseases. Curr. Med. Chem. 2013, 20, 3500–3514.
  39. Farokhzad, O.C.; Langer, R. Impact of nanotechnology on drug delivery. ACS Nano 2009, 3, 16–20.
  40. Marrache, S.; Dhar, S. Engineering of blended nanoparticle platform for delivery of mitochondriaacting therapeutics. Proc. Natl. Acad. Sci. USA 2012, 109, 16288–16293.
  41. Bae, Y.; Jung, M.K.; Lee, S.; Song, S.J.; Mun, J.Y.; Green, E.S.; Han, J.; Ko, K.S.; Choi, J.S. Dequalinium-based functional nanosomes show increased mitochondria targeting and anticancer effect. Eur. J. Pharm. Biopharm. 2018, 124, 104–115.
  42. Vaidya, B.; Paliwal, R.; Rai, S.; Khatri, K.; Goyal, A.K.; Mishra, N.; Vyas, S.P. Cell-selective mitochondrial targeting: A new approach for cancer therapy. Cancer Ther. 2009, 7, 141–148.
  43. Choi, Y.S.; Cho, T.S.; Kim, J.M.; Han, S.W.; Kim, S.K. Amine terminated G-6 PAMAM dendrimer and its interaction with DNA probed by Hoechst 33258. Biophys. Chem. 2006, 121, 142–149.
  44. Yu, Y.; Wang, Z.H.; Zhang, L.; Yao, H.J.; Zhang, Y.; Li, R.J.; Ju, R.J.; Wang, X.X.; Zhou, J.; Li, N.; et al. Mitochondrial targeting topotecan-loaded liposomes for treating drug-resistant breast cancer and inhibiting invasive metastases of melanoma. Biomaterials 2012, 33, 1808–1820.
  45. Wang, H.; Zhang, F.; Wen, H.; Shi, W.; Huang, Q.; Huang, Y.; Xie, J.; Li, P.; Chen, J.; Qin, L.; et al. Tumor- and mitochondria-targeted nanoparticles eradicate drug resistant lung cancer through mitochondrial pathway of apoptosis. J. Nanobiotechnol. 2020, 18, 8.
  46. Bonaccorso, A.; Pellitteri, R.; Ruozi, B.; Puglia, C.; Santonocito, D.; Pignatello, R.; Musumeci, T. Curcumin Loaded Polymeric vs. Lipid Nanoparticles: Antioxidant Effect on Normal and Hypoxic Olfactory Ensheathing Cells. Nanomaterials 2021, 11, 159.
  47. Sharma, A.; Liaw, K.; Sharma, R.; Zhang, Z.; Kannan, S.; Kannan, R.M. Targeting Mitochondrial Dysfunction and Oxidative Stress in Activated Microglia using Dendrimer-Based Therapeutics. Theranostics 2018, 8, 5529–5547.
  48. Song, Y.; Wang, X.; Wang, X.; Wang, J.; Hao, Q.; Hao, J.; Hou, X. Osthole-Loaded Nanoemulsion Enhances Brain Target in the Treatment of Alzheimer’s Disease via Intranasal Administration. Oxidative Med. Cell. Longev. 2021, 2021, 8844455.
  49. Elekofehinti, O.O.; Ayodele, O.C.; Iwaloye, O. Momordica charantia nanoparticles promote mitochondria biogenesis in the pancreas of diabetic-induced rats: Gene expression study. Egypt. J. Med. Hum. Genet. 2021, 22, 80.
  50. Weissig, V.; Lasch, J.; Erdos, G.; Meyer, H.W.; Rowe, T.C.; Hughes, J. DQAsomes: A novel potential drug and gene delivery system made from Dequalinium. Pharm. Res. 1998, 15, 334–337.
  51. Wang, L.; Liu, Y.; Li, W.; Jiang, X.; Ji, Y.; Wu, X.; Xu, L.; Qiu, Y.; Zhao, K.; Wei, T.; et al. Selective targeting of gold nanorods at the mitochondria of cancer cells: Implications for cancer therapy. Nano Lett. 2011, 11, 772–780.
  52. Kreyling, W.G.; Semmler-Behnke, M.; Seitz, J.; Scymczak, W.; Wenk, A.; Mayer, P.; Takenaka, S.; Oberdörster, G. Size dependence of the translocation of inhaled iridium and carbon nanoparticle aggregates from the lung of rats to the blood and secondary target organs. Inhal. Toxicol. 2009, 21, 55–60.
  53. Recordati, C.; De Maglie, M.; Bianchessi, S.; Argentiere, S.; Cella, C.; Mattiello, S.; Cubadda, F.; Aureli, F.; D’Amato, M.; Raggi, A.; et al. Tissue distribution and acute toxicity of silver after single intravenous administration in mice: Nano-specific and size-dependent effects. Part. Fibre Toxicol. 2016, 13, 12.
  54. Zoroddu, M.A.; Medici, S.; Ledda, A.; Nurchi, V.M.; Lachowicz, J.I.; Peana, M. Toxicity of nanoparticles. Curr. Med. Chem. 2014, 21, 3837–3853.
  55. Paek, H.J.; Lee, Y.J.; Chung, H.E.; Yoo, N.H.; Lee, J.A.; Kim, M.K.; Lee, J.K.; Jeong, J.; Choi, S.J. Modulation of the pharmacokinetics of zinc oxide nanoparticles and their fates in vivo. Nanoscale 2013, 5, 11416–11427.
More
Information
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: 353
Revisions: 2 times (View History)
Update Date: 03 Jul 2023
1000/1000
Video Production Service