Skip to main content

Horizontal mitochondrial transfer as a novel bioenergetic tool for mesenchymal stromal/stem cells: molecular mechanisms and therapeutic potential in a variety of diseases

Abstract

Intercellular mitochondrial transfer (MT) is a newly discovered form of cell-to-cell signalling involving the active incorporation of healthy mitochondria into stressed/injured recipient cells, contributing to the restoration of bioenergetic profile and cell viability, reduction of inflammatory processes and normalisation of calcium dynamics. Recent evidence has shown that MT can occur through multiple cellular structures and mechanisms: tunneling nanotubes (TNTs), via gap junctions (GJs), mediated by extracellular vesicles (EVs) and other mechanisms (cell fusion, mitochondrial extrusion and migrasome-mediated mitocytosis) and in different contexts, such as under physiological (tissue homeostasis and stemness maintenance) and pathological conditions (hypoxia, inflammation and cancer). As Mesenchimal Stromal/ Stem Cells (MSC)-mediated MT has emerged as a critical regulatory and restorative mechanism for cell and tissue regeneration and damage repair in recent years, its potential in stem cell therapy has received increasing attention. In particular, the potential therapeutic role of MSCs has been reported in several articles, suggesting that MSCs can enhance tissue repair after injury via MT and membrane vesicle release. For these reasons, in this review, we will discuss the different mechanisms of MSCs-mediated MT and therapeutic effects on different diseases such as neuronal, ischaemic, vascular and pulmonary diseases. Therefore, understanding the molecular and cellular mechanisms of MT and demonstrating its efficacy could be an important milestone that lays the foundation for future clinical trials.

Background

Mitochondria are highly dynamic and multifunctional compartments that play a pivotal role in oxidative bioenergetic metabolism and in maintaining cellular homeostasis. In addition to energy production, mitochondria perform many other key functions regulating fatty acid β-oxidation, iron homeostasis, Ca2+ metabolism, heme and steroid hormones biosynthesis, innate immunity, redox homeostasis, and cellular waste management [1,2,3]. They are key sensors of multiple types of cellular stress [4,5,6] and crucial determinants in cell survival and death [7]. To fulfill its multiple tasks, mitochondrial network is dynamically redesigned by distinct processes (collectively referred to as mitochondrial quality control, MQC), including fission and fusion events, intracellular mitochondrial movement, selective removal of damaged mitochondria through mitophagy [8,9,10], and mitochondrial biogenesis, that orchestrate the overall shape, size, distribution, and connectivity of mitochondria. As signaling platforms, mitochondria communicate extensively with other cellular compartments [11] and can operate outside their intracellular confines exchanging information between cells, and even across organ systems [12]. Interestingly, different forms of circulating mitochondria (e.g., wrapped Mitos, mitochondria transported in vesicles or in platelets; free Mitos, mitochondria without a protective vesicle membrane) with distinct effects on immune cells have been found in both blood and cerebrospinal fluid (CSF) [8, 13,14,15,16].

It is not surprising that the concomitant increase of mitochondrial DNA (mtDNA) mutations and reactive oxygen species (ROS) generation, during the aging process, exacerbate mitochondrial dysfunction and dysregulates MQC, contributing to the pathogenesis of multiple age-associated diseases [17, 18], recently termed as non‐communicable diseases (NCDs) [19].

In this view, mitochondria-targeted therapeutic approaches have been explored over the past decade as potential treatments for tissue revitalization and homeostasis.

The phenomenon of intercellular transport of mitochondria between mammalian cells, also known as horizontal transfer of mitochondria, has recently attracted a renewed attention from the scientific community, representing an intriguing reparative strategy [20, 21]. Intercellular mitochondrial transfer (MT) is a novel form of cell-to-cell signalling involving the active incorporation of healthy mitochondria into stressed/injured recipient cells [22], contributing to restore the bioenergetic profile (ATP and mitochondrial membrane potential) and cell viability, to increase the mtDNA content, as well as to reduce inflammatory processes and normalize calcium dynamics [23].

MT can occur via multiple distinct molecular mechanisms, including tunnelling nanotubes (TNTs), extracellular vesicles (EVs), gap junction channels (GJCs), and other non-traditional routes, such as cell fusion and mitochondrial extrusion [24, 25] (Fig. 1).

Fig. 1
figure 1

Different mechanisms and routes of intercellular mitochondrial transfer. Exposure to stress signals (e.g., dysfunctional mitochondria, mitoDAMPs, and NAD+) triggers the transfer of healthy mitochondria from donor cells to stressed/injured recipient cells via the activation of distinct signaling pathways and different routes, such as TNTs (1), EVs (2), and GJs (3). Key participants in the development of TNTs are F-Actin, microtubule, and intermediate filaments. Microtubule-based mitochondrial movement is mediated by Miro/TRAK complex attached to KIF5 kinesin/dynein molecular motors, whereas myosin motors are involved in mitochondrial transport based on actin. The NAD+/CD38/cADPR/Ca2+ pathway is involved in the release of EVs. In the presence of NAD+, CD38 can synthetize cADPR through its cyclase activity. As second messenger, cADPR triggers the release of the intracellular Ca2+ from the Endoplasmic reticulum by acting on Ryanodine Receptor (RyR). Therefore, the calcium-mediated activation of exocyst complex leads to the release of vesicles into the extracellular space. Importantly, the endocytosis of EVs may also occur via these mechanisms. Cx43-GJs are often located to one end of specific TNTs and the transcellular transfer of mitochondria may occur via gap junction internalization

The mitochondrial donation from mesenchymal stromal/stem cells (MSCs) has gained increasing attention in recent years [26], suggesting its potential in stem cell therapy [21]. MSCs-mediated MT is emerging as a critical regulatory mechanism for cell and tissue regeneration, and damage repair, where a remarkable restoration of cellular bioenergetics and a reduction in oxidative stress have been reported [26, 27]. MSCs, as reported by Viswanathan et al., are to be better referred as Mesenchymal Stromal/Stem Cells (MSCs) given the prevalence of stromal cells with respect to the smaller fraction of true "mesenchymal stem cells” (https://doi.org/10.1016/j.jcyt.2019.08.002). MSCs have the ability to maintain tissue homeostasis and renewal via regulating functional activity of parenchymal cells, including haematopoietic stem cells (HSCs) [28, 29]. They can be isolated and expanded in vitro from almost all tissues, including bone marrow (BM), dental pulp (DP), umbilical cord (UC), adipose tissue (AD), placenta, and other sources [30,31,32,33]. Moreover, they have pleiotropic ability to differentiate into mesodermal lineage cells, including adipocytes, chondrocytes and osteocytes, or ectodermal lineage cells, such as neuronal and neuroglial cells [34]. In physiological conditions, the original microenvironment characteristics of MSCs (intercellular matrix, other cell types, soluble factors and humoral components) dictate their cellular fate [35]. On the other hand, MSCs reciprocally interact with the microenvironment through distinct processes, including immunomodulation and tissue repair. Therefore, MSCs display remarkable tissue regenerative properties given their ability to receive distinct signals from the surrounding tissues (e.g., damage-associated molecular patterns (DAMPs) and ROS) and constantly set up intricate intercellular communication networks with stressed/damaged cells.

Mitochondrial exchange and its beneficial effects were firstly reported by Spees and colleagues [36] in A549 ρ° cells with mtDNA deletion or defects. Specifically, the acquisition of healthy mitochondria in A549 ρ° cells following active interaction with human MSCs resulted in increasing levels of oxygen consumption, membrane potential, and intracellular ATP. Subsequent investigations in different in vitro and in vivo models have revealed the possibility of using mitochondria derived from MSCs for therapeutic purposes (e.g., stroke, lung and acute respiratory disorders, brain injury, muscle sepsis, and cardiac diseases), thus representing a novel strategy to treat many NCDs. Features such as immune privilege, significant migratory capacity to injured sites, fine-tuned redox balance, and low oxidative damage levels, as well as low energy demand (mitochondria in MSCs are quiescent and exhibit low activity level) render MSCs as elective donor cells in delivering functional mitochondria to diseased cells [37,38,39,40,41,42]. Furthermore, MSCs isolated from different tissue types (BM; AD; DP; UC; Wharton’s jelly, WJ) exhibit distinct bioenergetic signatures that influence their MT capacity, with AD-MSCs and BM-MSCs showing higher MT levels to cardiomyocytes than DP-MSCs and WJ-MSCs [37].

On the other hand, it has also been shown that MSCs-mediated MT plays a critical role in pathological states. In this sense, recent evidence indicates a dark side of MSCs in increasing malignancy of cancer cells, supporting tumor microenvironment and cancer progression, and providing metabolic flexibility and chemotherapy resistance [37, 43,44,45].

In this review, we highlight the therapeutic potential of MT from MSCs in restoring the bioenergetic metabolism and cell functionality/viability into stressed/injured recipient cells. Therefore, we provide a comprehensive overview of the molecular signals triggering MT, as well as the major routes and mechanisms mediating MT. Finally, recent advances in MSCs regenerative properties through MT process in different NCDs are proposed.

Intercellular mitochondrial transfer via multiple cellular structures and mechanisms

MT has been demonstrated in various contexts under physiological (e.g., in tissue homeostasis and stemness maintenance) and pathological conditions (e.g., hypoxia, inflammation, and cancer). MT is a highly regulated multistep process that requires the spatiotemporal orchestration of many factors, including distinct triggering events/signals, molecular processes governing the formation of transfer machinery, and regulatory elements that modulate mitochondrial transfer speed and duration. In this section we will discuss the current state of knowledge regarding molecular stress signals and modes of MT, as well as signaling pathways regulating the different patterns of mitochondrial exchange.

Mitochondrial transfer trigger signals

Depending on different forms and activation states, mitochondria can affect the functions of neighbouring cells performing complex activities as mediators of regenerative and anti-inflammatory effects or trigger factors of inflammatory reactions [8, 46,47,48]. Multiple stress conditions, including oxygen–glucose deprivation [49, 50] drug-induced oxidative stress [51, 52] and inflammation [53, 54] generate harmed and fragmented extracellular mitochondria that trigger MT between cells. Therefore, the tissue regenerative properties of MSCs via the acquisition of active mitochondria by injured cells may be prompted by stress signals such as dysfunctional mitochondria and DAMPs of mitochondrial origin (mitoDAMPs), including ROS, DNA, cardiolipin, ATP, N-formyl peptides, transcription factor A mitochondria (TFAM), Cytochrome C (Cyt-C), succinate, and Ca2+(Fig. 1).

Consistent with this hypothesis, in an in vivo model of myocardial infarction, mitochondria released from damaged cells operate as potential DAMPs in MSCs, inducing the activation of eme-oxygenase-1 (HO-1) signaling pathway and mitochondrial biogenesis. These events enhance MT and consequently potentiate the rescue ability of MSCs to damaged cells [55]. In acute myeloid leukemia (AML) cells, NADPH oxidase-2 (NOX2)-derived ROS regulates MT from the BM-MSCs to the AML through TNTs [56]. In line with this, the generation of oxidative stress in HSCs triggers MT from BM-MSCs. Specifically, infection by Gram-negative bacteria drives MT from the BM-MSCs to HSCs through a ROS-dependent mechanism involving the activation of p53 and its downstream Akt/PI3K/mTOR pathway, and Connexin 43 (CX43) Gap Junctions [57, 58]. Of note, the p53-dependent activation of Akt/PI3K/mTOR pathway also leads to the overexpression of Tumor Necrosis Factor Alpha Induced Protein 2 (TNFαip2) and the formation of TNTs. As described above, MSCs also promotes the transfer of depolarized mitochondria to macrophages in response to ROS generation [58].

In addition to ROS, the Cyt-C released in stressed cells can also stimulate MT. Therefore, in UV-damaged PC12 cells at early stage of apoptosis loss of Cyt-C from injured mitochondria activates MT to healthy cells through TNTs formation, thereby leading to the recovery of apoptotic PC12 cells [59].

CD38 is a multifunctional transmembrane glycoprotein responsible for the biosynthesis of two calcium-mobilizing second messengers, cyclic ADP-ribose (cADPR) and nicotinic acid adenine dinucleotide phosphate. Interestingly, CD38 may also play a role in promoting MT in different cellular models. Therefore, by a calcium-dependent mechanism involving CD38 and cADPR signalling, astrocytes promote the mitochondrial exchange to nearby neurons via microvesicles (MVs), thereby contributing to activation of neuroprotective and neurorecovery mechanisms after ischemic stroke [60]. In this case, CD38 expression may be related to the excessive release of excitotoxic glutamate from ischemic neuron, suggesting for this neurotransmitter a potential role in promoting astrocytes-mediated MT. A subsequent study further demonstrated that CD38/cADPR signaling and mitochondrial Rho GTPases (Miro) 1 and 2 contribute to the MT between astrocytes, and from neuronal cells into astrocytes [61]. CD38 has been indicated to exert a critical function during bone formation [62]. Very recently, it has been demonstrated that CD38/cADPR signaling plays a crucial role in mediating the differentiation and maturation of osteoblasts and osteoprogenitor cells through stimulating the secretion of mitochondria and mitochondrial-derived EVs [63].

Tunnelling nanotubes (TNTs) as a major route for mitochondria exchange

Multiple in vitro and in vivo models have highlighted the elective role played by TNTs in MT. Specifically, MT can occur from MSCs to lung macrophages via TNTs, resulting in higher phagocytic activity of macrophage cells in a model of acute respiratory distress syndrome [64]. In an in vitro simulated ischemia/reperfusion model, MSCs move mitochondria to injured H9c2 cardiomyoblasts via TNTs, protecting cardiac cells against the apoptosis [65]. A bidirectional MT between MSCs and vascular smooth muscle cells via TNTs upregulates MSCs proliferation [66]. Also, MT via TNTs occurs from human BM-MSCs and endothelial cells experiencing chemotherapy stress, resulting in recovery of injured endothelial cells [67]. Mitochondria move between BM-MSCs and myeloma cells via TNTs, as described above [56, 68]. TNTs-mediated mitochondria exchange also form between donor cells other than MSCs, including PC12 cells [59], astrocytes [69], lung epithelial cells [70], cancer cells [71], retinal pigment epithelium [72], and trabecular meshwork cells [73].

Structural features of TNTs

According to the initial description, TNTs are open-ended membranous channels directly connecting cytoplasm of cells over long distances in a homotypic and heterotypic fashion. They are non-surface adherent and have a structure of 50–900 nm (with an average of 200 nm) in width and an average length between 20 and 100 mm.

TNTs contain F-actin cytoskeletal filaments that allow the bidirectional and unidirectional exchange between cells of various‐sized cargoes, including small molecules, nucleic acids and proteins (e.g., tau, α-synuclein, and huntingtin), organelles (e.g., vesicles, lysosomes, endoplasmic reticulum (ER), mitochondria, and autophagosomes), and even virus (herpesvirus and SARS-CoV-2) and bacteria [74,75,76,77,78], suggesting their role in in coordinating metabolism and signalling events in a wide-range of physiological processes and pathological conditions. Of note, this transfer function represents a crucial distinctive characteristic between TNTs and filopodia. The membrane structure of TNTs is very heterogeneous and cell type-specific. Therefore, close-ended TNTs (exhibiting gap junctions), that allow transfer of electrical signals, have also been identified in different in vitro and in vivo conditions [72, 79,80,81]. In this regard, 3 types of close-ended TNTs have been observed, such as “hand-shake”, “invaginated”, and “resting” [82].

In addition, two types of open-ended TNTs according to the different distribution of cytoskeletal elements have been described: “thin” TNTs (< 700 nm in diameter), containing only F-Actin, and “thick” TNTs (> 700 nm in diameter), with F-Actin, microtubule, and intermediate filaments [83]. In neuronal and stromal cell lines, cryo-correlative light and electron microscopy (cryo-CLEM) coupled with tomography have also revealed individual TNTs (iTNTs), a bundle of small open-ended tubes (up of two to 11) that run parallel, and sometimes braided together, along the entire length of the TNTs [82, 84]. These structures exhibit N-Cadherin as molecular linker connecting adjacent iTNTs. In addition to confer mechanical stability, it acts as guidance during growing iTNTs. Differently from single TNT, iTNTs allow for a bidirectional transfer of cargoes, including vesicles and mitochondria.

Mechanisms of TNT formation and protein regulators

Thus far, we have learned that the formation of TNTs can be generated via two different mechanisms that can occur simultaneously, or change dependently on the environmental conditions, though the underlying molecular process still remains under active investigation. These mechanisms include the “cell dislodgement”, wherein cells initially in contact leave behind a tubular connection when they move apart, and the “protrusion-elongation mechanism” where the cell extends a filopodia-like protrusion to another cell located at some distance. In this regard, the main mechanism responsible for TNTs biogenesis seems to be cell dislodgement, as reported by live imaging analyses in different cell types [85, 86]. However, the protrusion-elongation mechanism is typical of post-mitotic cells with low migratory phenotype, such as neurons and epithelial cells [87].

As mentioned above, the biogenesis of TNTs involves stress-signaling (e.g., p53 and MAP kinase) and pro-survival (e.g., EGFR, Akt, ROCK, PAK, MAP/ERK, PI3K or mTOR) pathways. Downstream effectors of these signaling pathways are proteins related to both membrane recycling and cytoskeletal remodeling.

Interestingly, the Rab family of small GTPases regulates many steps of membrane trafficking and also participate in actin cytoskeleton remodeling [88, 89]. Therefore, transport and recycling of vesicles regulated by the small GTPases Rab11a and Rab8a promote TNTs formation in different cell systems [88, 90]. Specifically, VAMP3 (vesicle-associated membrane protein 3, often called v-SNARES) operates downstream of Rab8a to regulate TNTs biogenesis [88]. Depletion of Ras association domain-containing protein 1A, a master regulator of cellular homeostasis and cytoskeleton, results in Rab11 accumulation and the subsequent release of exosome, thereby leading to TNTs formation [91]. Furthermore, Rab35 and its downstream effectors, such as ACAP2, ARF6-GDP, and EHD1 operate in a cascade mechanism to promote TNTs biogenesis in neuronal cells [92].

F-actin remodeling processes are responsible for initiating TNT protrusion by acting against the tensile strength of the plasma membrane, thus allowing it to deform and generate forces for tube growth. In this regard, distinct mechanisms involving specific sets of actin regulatory proteins are implicated, such as M-Sec (also known as TNFAIP2) [93,94,95] the exocyst complex, leukocyte specific transcript 1 (LST1) [96], the unconventional Myosin X (Myo10) [97], the actin bundler epidermal growth receptor substrate 8 (Eps8) and insulin receptor tyrosine kinase substrate protein 53 kDa (IRSp53), and small GTPases (e.g., Miro1/2, Rac1, Cdc42 and RalA) [98].

The cytosolic protein M-Sec acts as a key regulator of TNTs biogenesis through interaction with the small GTPase Ral in the mouse macrophage-like cell line RAW264.7 [93]. In this context, the transmembrane MHC class III protein LST1 may operate as a membrane scaffold for the generation of multi-molecular complex that controls the formation of TNTs. Thus, LST1 promotes TNTs formation by recruiting RalA and the actin-crosslinking protein filamin to the plasma membrane. It also stimulates the binding of RalA with two components of exocyst complex, Sec5 and Exo84, thereby leading to actin cytoskeletal remodeling and membrane protrusion. Meanwhile, the interaction of LST1 with M-Sec, myosin, and myoferlin may be involved in the process of mitochondrial anchoring and transfer [96]. In this regard, the recruitment of cytosolic M-Sec to the plasma membrane during the initial phase of TNT formation may occur through its direct binding to phosphatidylinositol (4,5)-bisphosphate [PI(4,5)P2] or phosphatidylinositol (3,4,5)-trisphosphate [PI(3,4,5)P3] [94]. Nucleolin, an RNA-binding protein, is essential for TNT formation in various types of mammalian cells [99]. Specifically, the binding of nucleolin to 14-3-3ζ mRNA modulates phospho-cofilin levels, leading to its inactivation, F-actin polymerization, and TNTs biogenesis.

In addition to the Rab35 signaling, Wingles-related integration site (Wnt)/Ca2+ pathway, an intracellular cascade that is implicated in actin cytoskeleton remodeling, promotes TNTs biogenesis and stability in CAD (mouse catecholaminergic neuronal cell line) cells and primary neurons through the modulation of the interaction between the β isoform of Ca2+/calmodulin‐dependent protein kinase II and actin [100].

The I-Bin/Amphiphysin/Rvs (BAR)-domain protein IRSp53 is an essential spatio-temporal coordinator of plasma membrane protrusions (through promoting negative membrane curvature) that couples Rho-GTPase signaling to cytoskeleton remodeling and membrane dynamics. As a signaling platform, IRSp53 operates under the control of activatory (Cdc42, Eps8) or inhibitory (14-3-3) inputs as well as downstream effectors, and recruits to the plasma membrane various actin regulatory proteins, such as Eps8, Ena/VASP4, Wiskott-Aldrich syndrome protein (WASP), N-WASP, mDia2, WASP family verprolin-homologous 2 (WAVE2), and others [101, 102].

Therefore, Delage and colleagues [98] have shown that Cdc42/IRSp53/VASP network negatively regulates TNTs formation and vesicular transport in neuronal cells, suggesting that Cdc42-dependent pathways are mainly involved in filopodia formation, rather than TNTs generation. By contrast, Eps8 (that reduces filopodia formation) is a positive regulator of TNTs biogenesis as its overexpression leads to increases in the extent of TNT connections and cargo transfer. Interestingly, the inhibition of actin related protein 2/3 complex (Arp2/3; a seven-subunit protein complex that mediates the formation of branched networks of filamentous actin, crucial for both filopodia and lamellipodia formation) promotes TNTs biogenesis and actin polymerization over longer distance in neuronal cells [82]. In particular, Arp2/3 inhibition enhances the co-expression of Eps8 and IRSp53, thus favoring their synergistic interaction. This event promotes the reorganization of the actin cytoskeleton, thereby leading to the switch of the system towards the extension of straight F-actin formation rather than the formation of branched networks [87] (Fig. 2). In addition, another recent study has demonstrated that the inhibition of ROCK, a downstream effector of Rho/Rac/Cdc42, stimulates TNTs formation via Myosin II mediated F-actin modulation [103].

Fig. 2
figure 2

Current model of TNTs formation via Eps8-IRSp53 interaction. In the cellular environment, the assembly of actin can be organized into two general types of architectures in balance with each other, such as branched actin networks and closely packed parallel arrays. The inhibition of Arp2/3-dependent branched filaments enhances the synergic activity of Eps8-IRSp53 complex at the membrane causing the shift of the actin balance in favour of polymerization of long, linear filaments for TNTs outgrowth. In this context, Eps8 seems to play a crucial role, possibly through renewed interactions with proteins and signaling pathways involved in actin dynamics, including phospho-regulators of cofilin (e.g., slingshot protein phosphatases and LIM-kinases), and coronins (e.g., Coro2b). Therefore, changes in Eps8’s interactome would lead to drastic reduction of actin filament turnover and disassembly of Arp2/3 networks, while promoting the extension and organization of actin

However, the same Rho GTPase signaling pathways operating in neuronal cells may act in a different way depending on the cell type. Indeed, observations in macrophages indicate that the blockage of Arp2/3 results in a decrease of TNTs formation [104]. Moreover, pathways converging on Arp2/3, including Cdc-42-WASP and Rac1-WAVE2 contribute together to TNT biogenesis. Consistent with this, in human trabecular meshwork cells (HTMCs; endothelial-like cells), the inactivation of Arp2/3 complex causes a reduction in TNT number and length, as well as vesicles transfer [73]. On the contrary, disassembly of actin stress fibers formation using Rho kinases inhibitors increased the number of TNTs and vesicle transfer.

Myo10 is an unconventional myosin that has critical functions in TNTs and filopodia formation [97, 105]. In particular, neuronal TNTs biogenesis requires both the N-terminal head (the myosin motor domain that can bind to F-actin) and C-terminal tail domains (that can bind to several molecules, such as PI(3,4,5)P3, microtubules and β-integrins) of Myo10. Myo10 operates downstream of Cdc42 and promotes TNTs formation independent of VASP (a strong inducer of dorsal filopodia) and through a mechanism independent of integrins and substrate attachment, suggesting that Myo10-driven TNTs may arise from a different mechanism than the one that drives the formation of dorsal filopodia [98, 105, 106].

In neurons, the extracellular protein S100A4 and its receptor RAGE (Receptor for Advanced Glycation End Product) play crucial role in directing TNTs biogenesis [107]. Under oxidative stress, p53-caspase-3 axis determines the cleavage and depletion of S100A4, thereby leading to a concentration gradient between neurons and astrocytes that drives the extension direction of TNTs from neurons (with low concentration of S100A4) to astrocytes (with high concentration of S100A4).

Role of Miro proteins in mitochondrial transfer via TNTs

Miro 1 and 2 proteins are embedded in the outer mitochondrial membrane and involved in the structure of the mitochondrial contact-site and cristae organizing system complex (that connects outer and inner-mitochondrial membrane), Ca2+ metabolism, mitochondria-ER communication, and mitophagy. Besides these functions, Miro proteins have critical role in regulating the mitochondrial spatial positioning and transport acting as adaptors that link mitochondria to cytoskeleton-associated motor proteins [108, 109]. Generally, microtubules and actin microfilament system mediate long- and short-distance transport of mitochondria, respectively. In this regard, movement of mitochondria based on microtubules involves the interaction of Miro/TRAK (Trafficking Kinesin Protein) complex with different motor proteins, such as Kinesin, dynein, and kinesin superfamily KIF5 [110, 111], while the actin-based mitochondrial movement is mediated by myosin family members, including myosin II, V, VI, and XIX [112, 113].

Recent studies demonstrate the essential role of Miro proteins in TNTs‐mediated mitochondrial [38] transport [114,115,116,117,118]. Specifically, overexpression of Miro1 promotes MT from donor MSCs to epithelial injured cells, leading to the rescue of epithelial functions. On the contrary, Miro1 knockdown inhibits MSC-mediated MT [114]. Consistent with this, upregulation of Miro1 in MSCs increases the metabolic/bioenergetic benefits of MT following neuronal oxidative stress and mitochondrial damage, while decreasing Miro1 expression reduces these effects [115, 117]. Also, MT via TNTs from astrocytes to neurons rescues neurons from cisplatin-induced damage. Furthermore, siRNA-mediated knockdown of Miro1 in astrocytes reduces MT, thus preventing the normalization of neuronal calcium dynamics [116].

Mitochondrial transfer via gap junctions (GJs)

MT can also occur in a Cx43-dependent manner since the deletion of Cx43 in donor/acceptor cells usually negatively affects the process [119,120,121]. Canonically, the formation of GJs or hemichannels across the plasma membrane by six Cx43 monomers allows uni- and bidirectional transfer of ions, small molecules (e.g., glucose, prostaglandins, microRNAs, and secondary messengers), and organelles between cells, thereby regulating intracellular mechanisms of signaling and several cellular functions [50, 122]. In line with this context, the MT from BM-MSCs to injured alveolar epithelium requires the Ca2+ exchanges between the two cells via Cx43-GJs, in the lipopolysaccharide (LPS)-induced acute lung injury mouse model [119]. In addition, MT from hematopoietic progenitors to BM-MSCs is required to induce the metabolic recovery of recipient BM-MSCs following irradiation and is cell-contact dependent and mediated by Cx43 [121].

Besides this classical function, Cx43-GJs may play further physiological roles, such as the regulation of mitochondrial functions and mediation of MT through the formation of TNTs and GJs internalization. In particular, Cx43 has also been detected within mitochondria [123] and seems to play a crucial role in mitochondrial calcium homeostasis and cell survival [124]. Cx43 is also implicated in forming connections between TNTs [125, 126] and marked reduction of TNTs is observed following knocking out the Cx43 genes in human trabecular meshwork [127]. Also, Cx43-mediated TNTs formation is crucial for MT from human induced pluripotent stem cell (iPSC)-derived MSCs to the injured bronchial epithelial cells and results in the inhibition of asthma inflammation [120]. Importantly, a novel mechanism for MT mediated by Cx43-GJs has been reported [128]. Specifically, this process requires a distinct form of GJs turnover involving the engulfment of GJs (a form of trogocytosis). Thus, following GJs internalization whole mitochondria and endosomes are incorporated into vesicles (named connexosomes/annular gap junctions) and transferred between neighbouring granulosa cells.

Mitochondrial transfer mediated by extracellular vesicles (EVs)

EVs are heterogeneous, phospholipid-enclosed structures that play a pivotal role in cell-to-cell communication at a longer range [129]. They are released by any cells into the extracellular space under different physiological and pathological conditions and encompass small EVs or exosomes (diameter of 30–200 nm), microvesicles (MVs or ectosomes; diameter of 100–1000 nm), and apoptotic bodies or apoptosomes (> 1 μm in size) [130, 131] Very recently, a specific subset of small EVs of mitochondrial origin has been identified in mouse and human brains named “mitovesicles” [132,133,134]. Mitochondria encapsulated into EVs can be transferred between cells to maintain the survival of metabolically compromised cells [58], to regulate immune responses [135], or to maintain tissue homeostasis [136]. In a mouse model of focal cerebral ischemia, MT by EVs between astrocytes and neurons acts as a survival mechanism protecting neurons from glucose deprivation and hypoxia [60]. EVs-mediated MT from renal scattered tubular cells (STC-like cells) to injured tubular epithelial cells (TEC) exerts protective effects, resulting in attenuating renal stenosis and recovering mitochondrial respiration [137]. Very recently, it has been demonstrated that mitochondria containing MVs (diameter of about 185 nm) derived from a human brain microvascular endothelial cell line significantly increase ATP levels, mitochondrial respiration, and glycolytic capacities of the ischemic primary human brain microvascular endothelial cells [138]. Interestingly, EVs-based interorgan transport of mitochondria may protect the heart through hormetic responses [139]. Specifically, circulating EVs derived from energetically stressed adipocytes, and containing oxidatively damaged mitochondria are taken up by cardiomyocytes, where they promote oxidative stress, resulting in a metabolic and redox adaptation of the heart that may confer protection against a future lethal stress.

In in vitro models of acute respiratory distress syndrome (ARDS), EVs-mediated MT from MSCs to human macrophages promotes phagocytosis and abolishes proinflammatory cytokine secretion [e.g., TNF-α] by macrophages, thereby mitigating lung injury [140]. Furthermore, in a model of allergic airway disease, myeloid-derived regulatory cells (MDRCs) encapsulate and transfer mitochondria to peripheral T cells via EVs, thereby affecting their bioenergetic and/or redox profile [141, 142]. Active mitochondria (in EVs and isolated) released by platelets stimulates the pro-angiogenic activity of MSCs via metabolic remodeling, including increased de novo fatty acid synthesis [143]. In this context, the clathrin-dependent endocytosis is the mechanism by which mitochondria are internalized by acceptor cells. Accordingly, EVs from MSCs are emerging as a novel nano-strategy approach to attenuate mitochondrial damage and improving TFAM-mtDNA complex stability, essential for regenerative capability of different tissues [144].

Other routes of mitochondrial transfer: cell fusion, mitochondrial extrusion, and migrasome-mediated mitocytosis

Although most data to date suggests that MT can occur via TNTs, Cx43-GJs, and EVs, various other routes have been proposed, including cell fusion, mitochondrial extrusion, and the migrasome-mediated mitocytosis [21, 145].

Cell fusion involves the physical merging of two or more cellular membranes, which would hypothetically allow the exchange of multiple protein complexes and even organelles, such as mitochondria [146]. Although cell fusion is infrequent under normal conditions, it may result in a marked mitochondrial delivery into recipient cells following injury and inflammation [147], hypoxia-induced apoptosis [148], or irradiation [149]. Cell fusion between human MSCs and cardiomyocytes results in MT into cardiomyocytes, and the resulting hybrid cells are reprogrammed toward a progenitor-like state [150]. Stem cells can fuse with other cellular models, including hepatocytes [151] and neurons [152], generating hybrid phenotypes which summarize distinct characteristics of both cells [153].

MT can also involve the extrusion or internalization of free mitochondria or mitochondrial components without membranous carriers. Under stress conditions, these events are crucial for regulating mitochondrial turnover and homeostasis [145, 154, 155]. Uncouplers of oxidative phosphorylation strongly stimulate the complete release of fragmented mitochondria from highly glycolytic HeLa cells via a mitoptosis process [154, 156, 157]. An analogous mechanism is also confirmed in in vivo models, including cardiomyocytes [158] and neurons [159, 160] suggesting a possible role as “waste removal” [145]. In Caenorhabditis elegans, under proteotoxic stress, neurons release dysfunctional mitochondria and protein aggregates via large membrane‐bound vesicles called “exophers” to support cellular homeostasis and functionality [160]. Murine cardiomyocytes can remove mitochondria and a significant portion of subcellular components via exopher‐like structures that are finally taken up and eliminated by cardiac macrophages. This process would support heart homeostasis by preventing extracellular accumulation of waste material, autophagic block, and inflammasome activation [158].

In hepatocytes and fibroblasts, the extrusion of mitochondria is also promoted by LPS through a process resembling secretory autophagy [161]. TNF-α induces the release of naked mitochondria into the extracellular spaces [155]. Similarly, massive mitochondrial extrusion has been shown in cell models of TNF‐α-induced necroptosis [162]. In this regard, free mitochondria would act as a specific danger signal to trigger inflammatory processes. Therefore, activated platelets can also release active mitochondria, both within membranous carriers and as free compartments, able to induce inflammation [163].

Importantly, MT may also occur via migrasomes, vesicular structures that grow on the tips and intersections of retraction fibers of migrating cells via tetraspanin microdomains [164, 165]. Therefore, to maintain cellular homeostasis, damaged mitochondria would be transported to the cell periphery before disposal by mitocytosis, an important mitochondrial quality-control process [166]. This process would deliver damaged mitochondria to surrounding cells [167].

Mesenchymal stem cells-mediated therapeutic effects of MT on different diseases

MT-mediated by MSCs may be used in restoring the bioenergetic metabolism and cell functionality and may be a useful tool for several diseases’ treatment. The MSCs potential therapeutic role has reported in several papers with a proposed mechanism which suggests that MSCs may enhance tissue repair after injury by MT and shedding of membrane vesicles [168]. MT is a strategy investigated in several kinds of cells such as pulmonary, cardiac, renal, corneal epithelium, and brain cortical [37]. Rustom et al. reported mitochondrial donation via a new form of cell-to-cell interaction based on TNTs [169]. Moreover, Jang et al. described intercellular mitochondrial transportation from MSCs to corneal endothelial cells, photoreceptors, and retinal pigment cells [170] and that recipient cells exhibited increased mitochondrial respiratory abilities. In the same paper Jang et al. reported that direct contact is a prerequisite for TNT formation and identified F-actin-based TNTs bridging MSCs and recipient cells. Several authors reported that mitochondrial dysfunction is associated also with several neurological diseases such as stroke, spinal cord injury and Alzheimer diseases (AD) [171, 172]. Moreover, there are evidence that MT may be a new approach to restoring mitochondrial functions and that the MT can be used to correct a range of problems caused by mitochondrial dysfunction [173, 174].

MT transfer and neuronal diseases

In several retinal and corneal diseases has been shown that mitochondrial dysfunction is a critical phenomenon [92, 175]. For example, it has observed that the increase of mitochondrial fission and mitochondrial DNA damage in retinal vasculature precede apoptosis of retinal endothelial cells in diabetic retinopathy [176, 177]. Therefore, several authors thought that targeting mitochondrial dysfunction may be an approach to prevent the development and progression of both retinal and corneal degeneration. Jiang et al. showed that intercellular mitochondrial transport is a vital mechanism for regeneration of corneal epithelial cells and retinal ganglion cells [51, 178]. Moreover, Jiang et al. describe intercellular mitochondrial transportation from MSCs to corneal endothelial cells, photoreceptors, and retinal pigment cells. In this paper they showed that the cell that receive the mitochondrion increased respiratory abilities and elevated expression of mitochondrial structure and function related gene [170]. Mitochondrial dysfunction has observed in Leber’s hereditary optic neuropathy (LHON). Recent evidence showed that LHON derives from a genetic mutation in mitochondria, which evolves to optic atrophy, which gives rise to visual acuity and blindness. Thus, the replacement of mitochondria may be a possible solution, which can be achieved by MT utilizing mesenchymal stem cells or their conditioned media derivate [179]. Several authors investigated a possible role of MT in AD.

AD is a chronic neurodegenerative disease and manifests symptoms such as: short-term memory loss, visual-spatial perception disorders and impaired language and executive functions [180]. Several papers showed that MSCs could inhibit amyloid β-peptide (Aβ) generation and promote its effective clearance, alter amyloid precursor protein (APP) processing, decrease tau phosphorylation, and increase proteasomal activity resulting in reduced accumulation of ubiquitin-conjugated proteins [181,182,183]. In AD has been observed the presence of Aβ in the mitochondria. This presence of Aβ generates hyperphosphorylated tau proteins by interacting with mitochondrial Drp1 protein, which, in turn, disrupts microtubule function and induces neural toxicity [184, 185]. Moreover, mitochondrial Aβ can also interact with Aβ-binding alcohol dehydrogenase (ABAD), leading to mitochondrial dysfunction and the production of ROS [186]. The main consequence of mitochondrial dysfunction, ROS accumulation, and increased oxidative stress are main factors involved in AD pathogenesis. Several studies showed that MSCs may promote microglia and autophagy-mediated clearance of protein aggregates as Aβ [187,188,189]. MSCs can protect neurons from cell death by secretion of some neuroprotective factors or by MT [52, 190, 191]. Zhan et al. showed that UC-MSC-CM significantly decreased tau phosphorylated at the Thr181 level, which increased in AD-alleviated intracellular and mitochondrial oxidative stress of okadaic acid (OA)-treated SH-SY5Y cells. In addition, UC-MSC-CM suppressed apoptosis and improved mitochondrial function in OA-treated SH-SY5Y cells. In this paper, Zhan et al. showed that UC-MSC-CM exerted protective effects relying on or partly extracellular vesicle (EV) MT from UC-MSCs to OA-treated SH-SY5Y cells [191]. Moreover, it observed that UC-MSC-CM decreased the level of p181-tau in the AD cell model, improved cell viability, and suppressed apoptosis in OA-treated Sh-SY5Y cells. In addition, they showed that UC-MSC-CM improved mitochondrial functions in OA-treated SH-SY5Y cells [192].

Spinal cord injury (SCI) is a destructive neurological disease that causes major motor, sensory and autonomic dysfunction that in some cases can lead blood vessels rupture and vasoconstriction reflexive led to an oxygen reduction and consequent mitochondrial damage. In addition, it is possible to observe a series of damages such as mitochondrial permeability, calcium overload, excitatory toxicity, oxidative stress, and increased ROS production [193]. These phenomenon results in impaired capability to maintain mitochondrial homeostasis with less energy available [194]. In this context, several authors are evolving new strategies to improve secondary injuries such as repairing or replacing damaged mitochondria, the use of antioxidants, and restoring mitochondrial permeability [195]. Li et al. showed that either MSCs or MSC-derived mitochondria injected into the injured spinal cord of a rat contusion SCI model significantly improved locomotor functions 6 weeks after injury [50]. Other authors reported that MSCs are able to improve the secondary injury caused by inflammation, myelin insulation, and assist the angiogenesis process [9, 196,197,198,199]. Sykova et al. showed the safety of the use of intravenous and intraarterial delivery of MSCs in SCI patients [200], while Deng et al. used MSCs coupled with collagen in SCI patients and compared with the control group (only collagen) [201]. After 12 months the study showed the treatment group vs control group showed significantly improved American Spinal Injury Association scores and better bowel and urinary functions [44].The ability of cells to interact with other cells via mitochondria has been demonstrated also in other tissues and organs where different modes of MT from MSCs to injured or damaged cells in order to restore or support nonfunctional mitochondria has been discovered [168]. For example, the active transfer from adult stem cells and somatic cells can rescue aerobic respiration in mammalian cells with non-functional mitochondria [36, 168].

MT transfer and ischemic vascular diseases

Another pathology that may benefit from studies on stem cell-mediated MT is stroke and ischemia–reperfusion injury. A stroke occurs when something blocks the blood supply to part of the brain or when a blood vessel in the brain bursts. Therefore, we can identify two types of strokes: ischemic or hemorrhagic. The blockage of one or more arteries is the main characteristic of acute ischemic stroke, with blood flow reduction and cellular dysfunction, damage, and/or death. The revascularization process is necessary for stroke treatment, but the oxygen and nutrient trasport to the damaged tissues may lead to the activation of the innate and adaptive immune responses that may cause secondary damage to the remaining cells [202, 203]. The hallmark of ischemia/reperfusion process is a mitochondrial dysfunction, ATP production decrease, ROS increase and cellular death [204]. During this process, cell switchs to anaerobic metabolism with a consequent indirect increase of Ca2+ [205]. This Ca2+ overload and oxidative stress lead to the opening of mitochondrial permeability transition pore in the inner mitochondrial membrane and an increase in ROS production [206, 207]. Therefore, MT from other cells could represent a useful tool in management of pathological damage caused by mitochondrial dysfunction in ischemia–reperfusion injury. It has been shown that several brain cells as neurons, astrocytes, endothelial cells, and MSCs are able to transfer mitochondria [171]. In preclinical studies, Liu et al. showed that human bone marrow MSCs could save endothelial cells during hypoxia and nutrients deprivation-induced stress. MSCs could abolish apoptosis in the endothelial cells induced by dysfunctional mitochondria during hypoxia by shifting the functional mitochondria from MSCs through TNTs-like cell protrusions [204]. Babenko et al. showed that BM-MSCs are able to save astrocytes and PC12 cells during hypoxia and glucose deprivation induced by oxidative stress and mitochondrial damage. Moreover, the same author reported that in the same in vivo study BM-MSCs improve the neurological impairments of cerebral ischemia rats and that this phenomenon is mediated by MT from BM-MSCs via TNTs. Moreover, it has been reported that BM-MSCs with overexpression of Miro1 improve the recovery of ischemic rats [115, 208]. Transplantation of healthy mitochondria has been studied as a solution in rescuing injured cells and tissue for treatment of others different pathologies connected to ischemia vascular diseases. The first studies have been conducted in animal models of ischemia–reperfusion [209]. For example, Mc Cully et al. have been demonstrated that administering mitochondria isolated from the left ventricle of the rabbit to the site of partial ischemia–reperfusion allowed a significant reduction of myocardial infarction and apoptosis markers with a consequent recovery of myocardial infarction [210]. Then, Masuzawa et al. demonstrated in their work the internalization of the transplanted mitochondria and enhanced myocardial energetics despite the lack of demonstration of internalization (by tunneling nanotubes) mechanism [211]. Similar results have been obtained in rat models of ischemia reperfusion. Kaza et al., administering mitochondria in in vivo rat models prior to reperfusion, obtained a decrease of Infarct Size and Area at Risk (IS/AAR) index and altogether enhanced myocardia and cell viability [212]. Similarly, Guariento and coworkers and Blitzer et al. investigated on the therapeutic use single or multiple intracoronary doses of isolated mitochondria before ischemia–reperfusion episodes [213, 214]. The first clinical application of mitochondrial transplantation was carried out to treat myocardial ischemia–reperfusion injury in pediatric patients of Boston Children’s Hospital (United States) [215]. The role of MT has been demonstrated also in SENECA trials where induced patients derived cardiomyocytes (iCM) co-cultured with MSCs, thanks to MSCs EV release, improved their viability and physiology, reducing ROS production and preserving mitochondrial biogenesis. Moreover, some authors suggested a mechanism by which EV from MSCs resulted enriched in mitochondria that probably were transferred to iCM [216]). Also Liu and colleagued have been demonstrated a mechanism by which MSCs can be used as a novel treatment of ischemic vascular disease; they demonstrated a transfer of mitochondria via a tunnelling nanotube-like structure from stem cells to injured human umbilical vein endothelial cells [204].

MT transfer and respiratory and pulmonary diseases

The effect of dysfunctional mitochondria has been reported also in respiratory and pulmonary diseases. Among these, Chronic Obstructive Pulmonary Disease (COPD) induced by cigarette smoke (CS) is an example of lung diseases characterized by inflammation and damage of cells. In particular, cigarette smoke induces mitochondrial disfunctions in lung epithelial cells. MSCs and MSC-derived exosomes have been proposed as therapeutic intervention of COPD induced by cigarette smoke as reported by Maremanda et al. which demonstrated as BEAS2B-mMSC co-cultures showed protective response against the CSE-altered mitochondrial respiration parameters, confirming the beneficial effect of MSC towards human bronchial lung epithelial cells [217]. Morrison et al. studying the mechanisms of MSCs effects in ARDS demonstrated that MSCs are able to induce in this environment an anti-inflammatory and highly phagocytic macrophage phenotype through EV-mediated MT. In particular, alveolar macrophage treated with MSC-derived EVs ameliorate lung injury in vivo. Transfer of functional mitochondria via EVs determined an increase in Oxidative phosphorylation which led to enhanced phagocytosis and a decrease of TNF-a and IL-8 levels of secretion by macrophages in vitro and in vivo [140]. Similarly, Jackson et al. reported that human bone marrow derived MSCs transfer their mitochondria to macrophages both in vivo and in vitro via TNT and micro vesicle secretion. This leads to enhanced macrophage phagocytosis and improved bioenergetics. So mitochondrial donation represented a novel mechanism to explain the antimicrobial effect of MSCs in a condition determined by bacterial infection providing additional evidence about their therapeutic use in acute, inflammatory lung diseases [64]. Moreover, the effect of mitochondrial transplantation or donation has been reported in other cell pathologies including renal and liver diseases. For example, Kubat et al. observed good therapeutic effects of mitochondrial transplantation in a nephrotoxicity model [218]. Otherwise, Lu et al. demonstrated a nanotherapeutic effect of UC-MSC-EVs on inhibiting local NETs formation by transferring functional mitochondria to intrahepatic neutrophils and repairing their mitochondrial function [219]. In addition, Bi et al. reported a functional therapeutic effect of mitochondrial transfer to combat Non-alcoholic fatty liver disease [220]. By the way, MT has been observed in lung diseases also associated with bioenergetic impairment and dysfunctional mitochondria such as allergic airway inflammation, ARDS or asthma but what signalling events can trigger this mechanism remain elusive.

Conclusions

Mitochondrial transfer and the mechanisms by which mitochondria enter recipient cells are concepts that have gained much attention in recent years. In particular, intercellular communication and mitochondrial transfer in MSCs promise interesting therapeutic results in various pathological conditions. In general, the active transfer of mitochondria from mesenchymal cells to somatic cells could, for instance, restore aerobic respiration in cells with non-functional mitochondria. So, the possible transplantation of healthy mitochondria has been studied as a solution in rescuing injured cells and tissue for treatment of several pathologies such as pulmonary, cardiac, renal, corneal epithelium, and brain cortical. Therefore, understanding the molecular and cellular mechanisms of mitochondrial transfer/transplantation and demonstrating its efficacy could be an important milestone that lays the foundation for future clinical trials.

Availability of data and materials

Not applicable.

Abbreviations

Arp2/3:

Actin related protein 2/3 complex

AML:

Acute myeloid leukemia

ARDS:

Acute respiratory distress syndrome

AD:

Adipose

AD:

Alzheimer diseases

BM:

Bone marrow

CSF:

Cerebrospinal fluid

COPD:

Chronic Obstructive Pulmonary Disease

CX43:

Connexin 43

Cyt-C:

Cytochrome C

DAMPs:

Damage-associated molecular patterns

DP:

Dental pulp

ER:

Endoplasmic reticulum

EVs:

Extracellular vesicles

GJs:

Gap junctions

HSCs:

Haematopoietic stem cells

iPSC:

Human induced pluripotent stem cell

HTMCs:

Human trabecular meshwork cells

IRSp53:

I-Bin/Amphiphysin/Rvs (BAR)-domain protein

iTNTs:

Individual TNTs

iCM:

Induced patients derived cardiomyocytes

LHON:

Leber’s hereditary optic neuropathy

MSCs:

Mesenchymal stem cells

MVs:

Microvesicles

MICOS:

Mitochondrial contact-site and cristae organizing system

mtDNA:

Mitochondrial DNA

MQC:

Mitochondrial quality control

MT:

Mitochondrial transfer

MDRCs:

Myeloid-derived regulatory cells

NOX2:

NADPH oxidase-2

NCDs:

Non‐communicable diseases

OMM:

Outer mitochondrial membrane

PI(4;5)P2 :

Phosphatidylinositol (4;5)-bisphosphate

ROS:

Reactive oxygen species

RAGE:

Receptor for Advanced Glycation End Product

SCI:

Spinal cord injury

TNFαip2:

TNF-α-induced protein 2

TFAM:

Transcription factor A mitochondria

TEC:

Tubular epithelial cells

TNF:

Tumor necrosis factor

TNTs:

Tunnelling nanotubes

UC:

Umbilical cord

WAVE2:

WASP family verprolin-homologous 2

WJ:

Wharton’s jelly

References

  1. Spinelli JB, Haigis MC. The multifaceted contributions of mitochondria to cellular metabolism. Nat Cell Biol. 2018;20(7):745–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Pfanner N, Warscheid B, Wiedemann N. Mitochondrial proteins: from biogenesis to functional networks. Nat Rev Mol Cell Biol. 2019;20(5):267–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Chakrabarty RP, Chandel NS. Mitochondria as signaling organelles control mammalian stem cell fate. Cell Stem Cell. 2021;28(3):394–408.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Iorio R, Castellucci A, Rossi G, Cinque B, Cifone MG, Macchiarelli G, et al. Mancozeb affects mitochondrial activity, redox status and ATP production in mouse granulosa cells. Toxicol In Vitro. 2015;30(1):438–45.

    Article  CAS  PubMed  Google Scholar 

  5. Petricca S, Flati V, Celenza G, Di Gregorio J, Lizzi AR, Luzi C, et al. Tebuconazole and econazole act synergistically in mediating mitochondrial stress, energy imbalance, and sequential activation of autophagy and apoptosis in mouse sertoli TM4 cells: possible role of AMPK/ULK1 axis. Toxicol Sci. 2019;169(1):209–23.

    Article  CAS  PubMed  Google Scholar 

  6. Di Gregorio J, Petricca S, Iorio R, Toniato E, Flati V. Mitochondrial and metabolic alterations in cancer cells. Eur J Cell Biol. 2022;101(3): 151225.

    Article  PubMed  Google Scholar 

  7. Bock FJ, Tait SWG. Mitochondria as multifaceted regulators of cell death. Nat Rev Mol Cell Biol. 2020;21(2):85–100.

    Article  CAS  PubMed  Google Scholar 

  8. Miliotis S, Nicolalde B, Ortega M, Yepez J, Caicedo A. Forms of extracellular mitochondria and their impact in health. Mitochondrion. 2019;48:16–30.

    Article  CAS  PubMed  Google Scholar 

  9. Zhou M, Yu Y, Luo X, Wang J, Lan X, Liu P, et al. Myocardial ischemia-reperfusion injury: therapeutics from a mitochondria-centric perspective. Cardiology. 2021;146(6):781–92.

    Article  CAS  PubMed  Google Scholar 

  10. Dong Y, Zhuang XX, Wang YT, Tan J, Feng D, Li M, et al. Chemical mitophagy modulators: drug development strategies and novel regulatory mechanisms. Pharmacol Res. 2023;194: 106835.

    Article  CAS  PubMed  Google Scholar 

  11. Picard M, Shirihai OS. Mitochondrial signal transduction. Cell Metab. 2022;34(11):1620–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Aguirre-Gutierrez J, Berenguer E, Menor IO, Bauman D, Corral-Rivas JJ, Nava-Miranda MG, et al. Functional susceptibility of tropical forests to climate change. Nat Ecol Evol. 2022;6(7):878.

    Article  PubMed  Google Scholar 

  13. Al Amir Dache Z, Otandault A, Tanos R, Pastor B, Meddeb R, Sanchez C, et al. Blood contains circulating cell-free respiratory competent mitochondria. FASEB J. 2020;34(3):3616–30.

    Article  CAS  PubMed  Google Scholar 

  14. Song X, Hu W, Yu H, Wang H, Zhao Y, Korngold R, et al. Existence of circulating mitochondria in human and animal peripheral blood. Int J Mol Sci. 2020;21(6):2122.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Stephens OR, Grant D, Frimel M, Wanner N, Yin M, Willard B, et al. Characterization and origins of cell-free mitochondria in healthy murine and human blood. Mitochondrion. 2020;54:102–12.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Chou SH, Lan J, Esposito E, Ning M, Balaj L, Ji X, et al. Extracellular mitochondria in cerebrospinal fluid and neurological recovery after subarachnoid hemorrhage. Stroke. 2017;48(8):2231–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Amorim JA, Coppotelli G, Rolo AP, Palmeira CM, Ross JM, Sinclair DA. Mitochondrial and metabolic dysfunction in ageing and age-related diseases. Nat Rev Endocrinol. 2022;18(4):243–58.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Lima T, Li TY, Mottis A, Auwerx J. Pleiotropic effects of mitochondria in aging. Nature aging. 2022;2(3):199–213.

    Article  PubMed  Google Scholar 

  19. Lim SS, Vos T, Flaxman AD, Danaei G, Shibuya K, Adair-Rohani H, et al. A comparative risk assessment of burden of disease and injury attributable to 67 risk factors and risk factor clusters in 21 regions, 1990–2010: a systematic analysis for the Global Burden of Disease Study 2010. Lancet. 2012;380(9859):2224–60.

    Article  PubMed  PubMed Central  Google Scholar 

  20. Shanmughapriya S, Langford D, Natarajaseenivasan K. Inter and Intracellular mitochondrial trafficking in health and disease. Ageing Res Rev. 2020;62: 101128.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Liu D, Gao Y, Liu J, Huang Y, Yin J, Feng Y, et al. Intercellular mitochondrial transfer as a means of tissue revitalization. Signal Transduct Target Ther. 2021;6(1):65.

    Article  PubMed  PubMed Central  Google Scholar 

  22. Al Amir Dache Z, Thierry AR. Mitochondria-derived cell-to-cell communication. Cell Rep. 2023;42(7):112728.

    Article  CAS  PubMed  Google Scholar 

  23. Liu Z, Sun Y, Qi Z, Cao L, Ding S. Mitochondrial transfer/transplantation: an emerging therapeutic approach for multiple diseases. Cell Biosci. 2022;12(1):66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Tiash S, Brestoff JR, Crewe C. A guide to studying mitochondria transfer. Nat Cell Biol. 2023;25(11):1551–3.

    Article  CAS  PubMed  Google Scholar 

  25. Clemente-Suarez VJ, Martin-Rodriguez A, Yanez-Sepulveda R, Tornero-Aguilera JF. Mitochondrial transfer as a novel therapeutic approach in disease diagnosis and treatment. Int J Mol Sci. 2023;24(10):8848.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Boukelmoune N, Chiu GS, Kavelaars A, Heijnen CJ. Mitochondrial transfer from mesenchymal stem cells to neural stem cells protects against the neurotoxic effects of cisplatin. Acta Neuropathol Commun. 2018;6(1):139.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Rodriguez AM, Nakhle J, Griessinger E, Vignais ML. Intercellular mitochondria trafficking highlighting the dual role of mesenchymal stem cells as both sensors and rescuers of tissue injury. Cell Cycle. 2018;17(6):712–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Mendez-Ferrer S, Michurina TV, Ferraro F, Mazloom AR, Macarthur BD, Lira SA, et al. Mesenchymal and haematopoietic stem cells form a unique bone marrow niche. Nature. 2010;466(7308):829–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Kfoury Y, Scadden DT. Mesenchymal cell contributions to the stem cell niche. Cell Stem Cell. 2015;16(3):239–53.

    Article  CAS  PubMed  Google Scholar 

  30. Martellucci S, Santacroce C, Manganelli V, Santilli F, Piccoli L, Cassetta M, et al. Isolation, propagation, and prion protein expression during neuronal differentiation of human dental pulp stem cells. J Vis Exp JoVE. 2019;145:e59282.

    Google Scholar 

  31. Delle Monache S, Pulcini F, Santilli F, Martellucci S, Santacroce C, Fabrizi J, et al. Hypoxia induces DPSC differentiation versus a neurogenic phenotype by the paracrine mechanism. Biomedicines. 2022;10(5):1056.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Galipeau J, Sensebe L. Mesenchymal stromal cells: clinical challenges and therapeutic opportunities. Cell Stem Cell. 2018;22(6):824–33.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Pittenger MF, Discher DE, Peault BM, Phinney DG, Hare JM, Caplan AI. Mesenchymal stem cell perspective: cell biology to clinical progress. NPJ Regenerative medicine. 2019;4:22.

    Article  PubMed  PubMed Central  Google Scholar 

  34. Salehinejad P, Moshrefi M, Eslaminejad T. An overview on mesenchymal stem cells derived from extraembryonic tissues: supplement sources and isolation methods. Stem Cells Cloning Adv Appl. 2020;13:57–65.

    Google Scholar 

  35. Liu J, Gao J, Liang Z, Gao C, Niu Q, Wu F, et al. Mesenchymal stem cells and their microenvironment. Stem Cell Res Ther. 2022;13(1):429.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Spees JL, Olson SD, Whitney MJ, Prockop DJ. Mitochondrial transfer between cells can rescue aerobic respiration. Proc Natl Acad Sci U S A. 2006;103(5):1283–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Paliwal S, Chaudhuri R, Agrawal A, Mohanty S. Regenerative abilities of mesenchymal stem cells through mitochondrial transfer. J Biomed Sci. 2018;25(1):31.

    Article  PubMed  PubMed Central  Google Scholar 

  38. Yang J, Liu L, Oda Y, Wada K, Ago M, Matsuda S, et al. Extracellular vesicles and Cx43-gap junction channels are the main routes for mitochondrial transfer from ultra-purified mesenchymal stem cells, RECs. Int J Mol Sci. 2023;24(12):10294.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Zhang F, Zheng X, Zhao F, Li L, Ren Y, Li L, et al. TFAM-Mediated mitochondrial transfer of MSCs improved the permeability barrier in sepsis-associated acute lung injury. Apoptosis. 2023;28(7–8):1048–59.

    Article  CAS  PubMed  Google Scholar 

  40. Malekpour K, Hazrati A, Soudi S, Hashemi SM. Mechanisms behind therapeutic potentials of mesenchymal stem cell mitochondria transfer/delivery. J Control Release. 2023;354:755–69.

    Article  CAS  PubMed  Google Scholar 

  41. Fahey M, Bennett M, Thomas M, Montney K, Vivancos-Koopman I, Pugliese B, et al. Mesenchymal stromal cells donate mitochondria to articular chondrocytes exposed to mitochondrial, environmental, and mechanical stress. Sci Rep. 2022;12(1):21525.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Thomas MA, Fahey MJ, Pugliese BR, Irwin RM, Antonyak MA, Delco ML. Human mesenchymal stromal cells release functional mitochondria in extracellular vesicles. Front Bioeng Biotechnol. 2022;10: 870193.

    Article  PubMed  PubMed Central  Google Scholar 

  43. Aponte PM, Caicedo A. Stemness in cancer: stem cells, cancer stem cells, and their microenvironment. Stem Cells Int. 2017;2017:5619472.

    Article  PubMed  PubMed Central  Google Scholar 

  44. Chen YC, Gonzalez ME, Burman B, Zhao X, Anwar T, Tran M, et al. Mesenchymal stem/stromal cell engulfment reveals metastatic advantage in breast cancer. Cell Rep. 2019;27(13):3916-26 e5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Sundstrom T, Prestegarden L, Azuaje F, Aasen SN, Rosland GV, Varughese JK, et al. Inhibition of mitochondrial respiration prevents BRAF-mutant melanoma brain metastasis. Acta Neuropathol Commun. 2019;7(1):55.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Patergnani S, Bouhamida E, Leo S, Pinton P, Rimessi A. Mitochondrial oxidative stress and “mito-inflammation”: actors in the diseases. Biomedicines. 2021;9(2):216.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Garg M, Johri S, Chakraborty K. Immunomodulatory role of mitochondrial DAMPs: a missing link in pathology? FEBS J. 2023;290(18):4395–418.

    Article  CAS  PubMed  Google Scholar 

  48. Marchi S, Guilbaud E, Tait SWG, Yamazaki T, Galluzzi L. Mitochondrial control of inflammation. Nat Rev Immunol. 2023;23(3):159–73.

    Article  CAS  PubMed  Google Scholar 

  49. Hayakawa K, Chan SJ, Mandeville ET, Park JH, Bruzzese M, Montaner J, et al. Protective effects of endothelial progenitor cell-derived extracellular mitochondria in brain endothelium. Stem cells. 2018;36(9):1404–10.

    Article  CAS  PubMed  Google Scholar 

  50. Li H, Wang C, He T, Zhao T, Chen YY, Shen YL, et al. Mitochondrial transfer from bone marrow mesenchymal stem cells to motor neurons in spinal cord injury rats via gap junction. Theranostics. 2019;9(7):2017–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Jiang D, Gao F, Zhang Y, Wong DS, Li Q, Tse HF, et al. Mitochondrial transfer of mesenchymal stem cells effectively protects corneal epithelial cells from mitochondrial damage. Cell Death Dis. 2016;7(11): e2467.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Wang J, Liu X, Qiu Y, Shi Y, Cai J, Wang B, et al. Cell adhesion-mediated mitochondria transfer contributes to mesenchymal stem cell-induced chemoresistance on T cell acute lymphoblastic leukemia cells. J Hematol Oncol. 2018;11(1):11.

    Article  PubMed  PubMed Central  Google Scholar 

  53. Zhang Y, Yu Z, Jiang D, Liang X, Liao S, Zhang Z, et al. iPSC-MSCs with high intrinsic MIRO1 and sensitivity to TNF-alpha yield efficacious mitochondrial transfer to rescue anthracycline-induced cardiomyopathy. Stem cell reports. 2016;7(4):749–63.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Luz-Crawford P, Hernandez J, Djouad F, Luque-Campos N, Caicedo A, Carrere-Kremer S, et al. Mesenchymal stem cell repression of Th17 cells is triggered by mitochondrial transfer. Stem Cell Res Ther. 2019;10(1):232.

    Article  PubMed  PubMed Central  Google Scholar 

  55. Mahrouf-Yorgov M, Augeul L, Da Silva CC, Jourdan M, Rigolet M, Manin S, et al. Mesenchymal stem cells sense mitochondria released from damaged cells as danger signals to activate their rescue properties. Cell Death Differ. 2017;24(7):1224–38.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Marlein CR, Zaitseva L, Piddock RE, Robinson SD, Edwards DR, Shafat MS, et al. NADPH oxidase-2 derived superoxide drives mitochondrial transfer from bone marrow stromal cells to leukemic blasts. Blood. 2017;130(14):1649–60.

    Article  CAS  PubMed  Google Scholar 

  57. Mistry JJ, Marlein CR, Moore JA, Hellmich C, Wojtowicz EE, Smith JGW, et al. ROS-mediated PI3K activation drives mitochondrial transfer from stromal cells to hematopoietic stem cells in response to infection. Proc Natl Acad Sci USA. 2019;116(49):24610–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Phinney DG, Di Giuseppe M, Njah J, Sala E, Shiva S, St Croix CM, et al. Mesenchymal stem cells use extracellular vesicles to outsource mitophagy and shuttle microRNAs. Nat Commun. 2015;6:8472.

    Article  CAS  PubMed  Google Scholar 

  59. Wang X, Gerdes HH. Transfer of mitochondria via tunneling nanotubes rescues apoptotic PC12 cells. Cell Death Differ. 2015;22(7):1181–91.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  60. Hayakawa K, Esposito E, Wang X, Terasaki Y, Liu Y, Xing C, et al. Transfer of mitochondria from astrocytes to neurons after stroke. Nature. 2016;535(7613):551–5.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Gao L, Zhang Z, Lu J, Pei G. Mitochondria are dynamically transferring between human neural cells and alexander disease-associated GFAP mutations impair the astrocytic transfer. Front Cell Neurosci. 2019;13:316.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Moridera K, Uchida S, Tanaka S, Menuki K, Utsunomiya H, Yamaoka K, et al. Skeletal unloading reduces cluster of differentiation (CD) 38 expression in the bone marrow and osteoblasts of mice. J Orthopaed Sci. 2020;25(2):331–7.

    Article  Google Scholar 

  63. Suh J, Kim NK, Shim W, Lee SH, Kim HJ, Moon E, et al. Mitochondrial fragmentation and donut formation enhance mitochondrial secretion to promote osteogenesis. Cell Metab. 2023;35(2):345-60 e7.

    Article  CAS  PubMed  Google Scholar 

  64. Jackson MV, Morrison TJ, Doherty DF, McAuley DF, Matthay MA, Kissenpfennig A, et al. Mitochondrial transfer via tunneling nanotubes is an important mechanism by which mesenchymal stem cells enhance macrophage phagocytosis in the in vitro and in vivo models of ARDS. Stem cells. 2016;34(8):2210–23.

    Article  CAS  PubMed  Google Scholar 

  65. Han H, Hu J, Yan Q, Zhu J, Zhu Z, Chen Y, et al. Bone marrow-derived mesenchymal stem cells rescue injured H9c2 cells via transferring intact mitochondria through tunneling nanotubes in an in vitro simulated ischemia/reperfusion model. Mol Med Rep. 2016;13(2):1517–24.

    Article  CAS  PubMed  Google Scholar 

  66. Vallabhaneni KC, Haller H, Dumler I. Vascular smooth muscle cells initiate proliferation of mesenchymal stem cells by mitochondrial transfer via tunneling nanotubes. Stem Cells Dev. 2012;21(17):3104–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Feng Y, Zhu R, Shen J, Wu J, Lu W, Zhang J, et al. Human bone marrow mesenchymal stem cells rescue endothelial cells experiencing chemotherapy stress by mitochondrial transfer via tunneling nanotubes. Stem Cells Dev. 2019;28(10):674–82.

    Article  CAS  PubMed  Google Scholar 

  68. Marlein CR, Piddock RE, Mistry JJ, Zaitseva L, Hellmich C, Horton RH, et al. CD38-driven mitochondrial trafficking promotes bioenergetic plasticity in multiple myeloma. Can Res. 2019;79(9):2285–97.

    Article  CAS  Google Scholar 

  69. Rostami J, Holmqvist S, Lindstrom V, Sigvardson J, Westermark GT, Ingelsson M, et al. Human astrocytes transfer aggregated alpha-synuclein via tunneling nanotubes. J Neurosci. 2017;37(49):11835–53.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Kumar A, Kim JH, Ranjan P, Metcalfe MG, Cao W, Mishina M, et al. Influenza virus exploits tunneling nanotubes for cell-to-cell spread. Sci Rep. 2017;7:40360.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Lu J, Zheng X, Li F, Yu Y, Chen Z, Liu Z, et al. Tunneling nanotubes promote intercellular mitochondria transfer followed by increased invasiveness in bladder cancer cells. Oncotarget. 2017;8(9):15539–52.

    Article  PubMed  PubMed Central  Google Scholar 

  72. Chinnery HR, Keller KE. Tunneling nanotubes and the eye: intercellular communication and implications for ocular health and disease. Biomed Res Int. 2020;2020:7246785.

    Article  PubMed  PubMed Central  Google Scholar 

  73. Keller KE, Bradley JM, Sun YY, Yang YF, Acott TS. Tunneling nanotubes are novel cellular structures that communicate signals between trabecular meshwork cells. Invest Ophthalmol Vis Sci. 2017;58(12):5298–307.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  74. Li RF, Zhang W, Man QW, Zhao YF, Zhao Y. Tunneling nanotubes mediate intercellular communication between endothelial progenitor cells and osteoclast precursors. J Mol Histol. 2019;50(5):483–91.

    Article  CAS  PubMed  Google Scholar 

  75. Panasiuk M, Rychlowski M, Derewonko N, Bienkowska-Szewczyk K. Tunneling nanotubes as a novel route of cell-to-cell spread of herpesviruses. J Virol. 2018;92(10):10–1128.

    Article  Google Scholar 

  76. de Rooij B, Polak R, Stalpers F, Pieters R, den Boer ML. Tunneling nanotubes facilitate autophagosome transfer in the leukemic niche. Leukemia. 2017;31(7):1651–4.

    Article  PubMed  PubMed Central  Google Scholar 

  77. Tiwari V, Koganti R, Russell G, Sharma A, Shukla D. Role of tunneling nanotubes in viral infection, neurodegenerative disease, and cancer. Front Immunol. 2021;12: 680891.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Pepe A, Pietropaoli S, Vos M, Barba-Spaeth G, Zurzolo C. Tunneling nanotubes provide a route for SARS-CoV-2 spreading. Sci Adv. 2022;8(29):eabo0171.

    Article  PubMed  PubMed Central  Google Scholar 

  79. Wang X, Bukoreshtliev NV, Gerdes HH. Developing neurons form transient nanotubes facilitating electrical coupling and calcium signaling with distant astrocytes. PLoS ONE. 2012;7(10): e47429.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  80. Chinnery HR, Pearlman E, McMenamin PG. Cutting edge: membrane nanotubes in vivo: a feature of MHC class II+ cells in the mouse cornea. J Immunol. 2008;180(9):5779–83.

    Article  CAS  PubMed  Google Scholar 

  81. Alarcon-Martinez L, Villafranca-Baughman D, Quintero H, Kacerovsky JB, Dotigny F, Murai KK, et al. Interpericyte tunnelling nanotubes regulate neurovascular coupling. Nature. 2020;585(7823):91–5.

    Article  CAS  PubMed  Google Scholar 

  82. Sartori-Rupp A, Cordero Cervantes D, Pepe A, Gousset K, Delage E, Corroyer-Dulmont S, et al. Correlative cryo-electron microscopy reveals the structure of TNTs in neuronal cells. Nat Commun. 2019;10(1):342.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Resnik N, Erman A, Veranic P, Kreft ME. Triple labelling of actin filaments, intermediate filaments and microtubules for broad application in cell biology: uncovering the cytoskeletal composition in tunneling nanotubes. Histochem Cell Biol. 2019;152(4):311–7.

    Article  CAS  PubMed  Google Scholar 

  84. Kolba MD, Dudka W, Zareba-Koziol M, Kominek A, Ronchi P, Turos L, et al. Tunneling nanotube-mediated intercellular vesicle and protein transfer in the stroma-provided imatinib resistance in chronic myeloid leukemia cells. Cell Death Dis. 2019;10(11):817.

    Article  PubMed  PubMed Central  Google Scholar 

  85. Reichert D, Scheinpflug J, Karbanova J, Freund D, Bornhauser M, Corbeil D. Tunneling nanotubes mediate the transfer of stem cell marker CD133 between hematopoietic progenitor cells. Exp Hematol. 2016;44(11):1092-112 e2.

    Article  CAS  PubMed  Google Scholar 

  86. Saenz-de-Santa-Maria I, Bernardo-Castineira C, Enciso E, Garcia-Moreno I, Chiara JL, Suarez C, et al. Control of long-distance cell-to-cell communication and autophagosome transfer in squamous cell carcinoma via tunneling nanotubes. Oncotarget. 2017;8(13):20939–60.

    Article  PubMed  PubMed Central  Google Scholar 

  87. Henderson JM, Ljubojevic N, Chaze T, Castaneda D, Battistella A, Gianetto QG, et al. A balance between actin and Eps8/IRSp53 utilization in branched versus linear actin networks determines tunneling nanotube formation. BioRXiv. 2022.

  88. Zhu S, Bhat S, Syan S, Kuchitsu Y, Fukuda M, Zurzolo C. Rab11a-Rab8a cascade regulates the formation of tunneling nanotubes through vesicle recycling. J Cell Sci. 2018;131(19):jcs215889.

    Article  PubMed  Google Scholar 

  89. Kuhns S, Seixas C, Pestana S, Tavares B, Nogueira R, Jacinto R, et al. Rab35 controls cilium length, function and membrane composition. EMBO Rep. 2019;20(10): e47625.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Zhu H, Xue C, Xu X, Guo Y, Li X, Lu J, et al. Rab8a/Rab11a regulate intercellular communications between neural cells via tunneling nanotubes. Cell Death Dis. 2016;7(12): e2523.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Dubois F, Jean-Jacques B, Roberge H, Benard M, Galas L, Schapman D, et al. A role for RASSF1A in tunneling nanotube formation between cells through GEFH1/Rab11 pathway control. Cell Commun Signal. 2018;16(1):66.

    Article  PubMed  PubMed Central  Google Scholar 

  92. Bhat S, Ljubojevic N, Zhu S, Fukuda M, Echard A, Zurzolo C. Rab35 and its effectors promote formation of tunneling nanotubes in neuronal cells. Sci Rep. 2020;10(1):16803.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Hase K, Kimura S, Takatsu H, Ohmae M, Kawano S, Kitamura H, et al. M-Sec promotes membrane nanotube formation by interacting with Ral and the exocyst complex. Nat Cell Biol. 2009;11(12):1427–32.

    Article  CAS  PubMed  Google Scholar 

  94. Kimura S, Yamashita M, Yamakami-Kimura M, Sato Y, Yamagata A, Kobashigawa Y, et al. Distinct roles for the N- and C-terminal regions of M-Sec in plasma membrane deformation during tunneling nanotube formation. Sci Rep. 2016;6:33548.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Lotfi S, Nasser H, Noyori O, Hiyoshi M, Takeuchi H, Koyanagi Y, et al. M-Sec facilitates intercellular transmission of HIV-1 through multiple mechanisms. Retrovirology. 2020;17(1):20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Schiller C, Diakopoulos KN, Rohwedder I, Kremmer E, von Toerne C, Ueffing M, et al. LST1 promotes the assembly of a molecular machinery responsible for tunneling nanotube formation. J Cell Sci. 2013;126(Pt 3):767–77.

    CAS  PubMed  Google Scholar 

  97. Gousset K, Marzo L, Commere PH, Zurzolo C. Myo10 is a key regulator of TNT formation in neuronal cells. J Cell Sci. 2013;126(Pt 19):4424–35.

    Article  CAS  PubMed  Google Scholar 

  98. Delage E, Cervantes DC, Penard E, Schmitt C, Syan S, Disanza A, et al. Differential identity of Filopodia and Tunneling Nanotubes revealed by the opposite functions of actin regulatory complexes. Sci Rep. 2016;6:39632.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Dagar S, Pushpa K, Pathak D, Samaddar S, Saxena A, Banerjee S, et al. Nucleolin regulates 14-3-3zeta mRNA and promotes cofilin phosphorylation to induce tunneling nanotube formation. FASEB J. 2021;35(1): e21199.

    Article  CAS  PubMed  Google Scholar 

  100. Vargas JY, Loria F, Wu YJ, Cordova G, Nonaka T, Bellow S, et al. The Wnt/Ca(2+) pathway is involved in interneuronal communication mediated by tunneling nanotubes. EMBO J. 2019;38(23): e101230.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Kast DJ, Dominguez R. Mechanism of IRSp53 inhibition by 14-3-3. Nat Commun. 2019;10(1):483.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Bisi S, Marchesi S, Rizvi A, Carra D, Beznoussenko GV, Ferrara I, et al. IRSp53 controls plasma membrane shape and polarized transport at the nascent lumen in epithelial tubules. Nat Commun. 2020;11(1):3516.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Scheiblich H, Dansokho C, Mercan D, Schmidt SV, Bousset L, Wischhof L, et al. Microglia jointly degrade fibrillar alpha-synuclein cargo by distribution through tunneling nanotubes. Cell. 2021;184(20):5089-106 e21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Hanna SJ, McCoy-Simandle K, Miskolci V, Guo P, Cammer M, Hodgson L, et al. The role of Rho-GTPases and actin polymerization during macrophage tunneling nanotube biogenesis. Sci Rep. 2017;7(1):8547.

    Article  PubMed  PubMed Central  Google Scholar 

  105. Bohil AB, Robertson BW, Cheney RE. Myosin-X is a molecular motor that functions in filopodia formation. Proc Natl Acad Sci USA. 2006;103(33):12411–6.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Kerber ML, Cheney RE. Myosin-X: a MyTH-FERM myosin at the tips of filopodia. J Cell Sci. 2011;124(Pt 22):3733–41.

    Article  PubMed  PubMed Central  Google Scholar 

  107. Sun X, Wang Y, Zhang J, Tu J, Wang XJ, Su XD, et al. Tunneling-nanotube direction determination in neurons and astrocytes. Cell Death Dis. 2012;3(12): e438.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Eberhardt EL, Ludlam AV, Tan Z, Cianfrocco MA. Miro: a molecular switch at the center of mitochondrial regulation. Protein Sci. 2020;29(6):1269–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  109. Nahacka Z, Zobalova R, Dubisova M, Rohlena J, Neuzil J. Miro proteins connect mitochondrial function and intercellular transport. Crit Rev Biochem Mol Biol. 2021;56(4):401–25.

    Article  CAS  PubMed  Google Scholar 

  110. van Spronsen M, Mikhaylova M, Lipka J, Schlager MA, van den Heuvel DJ, Kuijpers M, et al. TRAK/Milton motor-adaptor proteins steer mitochondrial trafficking to axons and dendrites. Neuron. 2013;77(3):485–502.

    Article  PubMed  Google Scholar 

  111. Birsa N, Norkett R, Higgs N, Lopez-Domenech G, Kittler JT. Mitochondrial trafficking in neurons and the role of the Miro family of GTPase proteins. Biochem Soc Trans. 2013;41(6):1525–31.

    Article  CAS  PubMed  Google Scholar 

  112. Oeding SJ, Majstrowicz K, Hu XP, Schwarz V, Freitag A, Honnert U, et al. Identification of Miro1 and Miro2 as mitochondrial receptors for myosin XIX. J Cell Sci. 2018;131(17):jcs219469.

    Article  PubMed  Google Scholar 

  113. Lopez-Domenech G, Covill-Cooke C, Ivankovic D, Halff EF, Sheehan DF, Norkett R, et al. Miro proteins coordinate microtubule- and actin-dependent mitochondrial transport and distribution. EMBO J. 2018;37(3):321–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  114. Ahmad T, Mukherjee S, Pattnaik B, Kumar M, Singh S, Kumar M, et al. Miro1 regulates intercellular mitochondrial transport & enhances mesenchymal stem cell rescue efficacy. EMBO J. 2014;33(9):994–1010.

    CAS  PubMed  PubMed Central  Google Scholar 

  115. Babenko VA, Silachev DN, Popkov VA, Zorova LD, Pevzner IB, Plotnikov EY, et al. Miro1 enhances mitochondria transfer from multipotent mesenchymal stem cells (MMSC) to neural cells and improves the efficacy of cell recovery. Molecules. 2018;23(3):687.

    Article  PubMed  PubMed Central  Google Scholar 

  116. English K, Shepherd A, Uzor NE, Trinh R, Kavelaars A, Heijnen CJ. Astrocytes rescue neuronal health after cisplatin treatment through mitochondrial transfer. Acta Neuropathol Commun. 2020;8(1):36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  117. Tseng N, Lambie SC, Huynh CQ, Sanford B, Patel M, Herson PS, et al. Mitochondrial transfer from mesenchymal stem cells improves neuronal metabolism after oxidant injury in vitro: The role of Miro1. J Cereb Blood Flow Metab. 2021;41(4):761–70.

    Article  CAS  PubMed  Google Scholar 

  118. Wang F, Chen X, Cheng H, Song L, Liu J, Caplan S, et al. MICAL2PV suppresses the formation of tunneling nanotubes and modulates mitochondrial trafficking. EMBO Rep. 2021;22(7): e52006.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Islam MN, Das SR, Emin MT, Wei M, Sun L, Westphalen K, et al. Mitochondrial transfer from bone-marrow-derived stromal cells to pulmonary alveoli protects against acute lung injury. Nat Med. 2012;18(5):759–65.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Yao Y, Fan XL, Jiang D, Zhang Y, Li X, Xu ZB, et al. Connexin 43-mediated mitochondrial transfer of iPSC-MSCs alleviates asthma inflammation. Stem cell reports. 2018;11(5):1120–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Golan K, Singh AK, Kollet O, Bertagna M, Althoff MJ, Khatib-Massalha E, et al. Bone marrow regeneration requires mitochondrial transfer from donor Cx43-expressing hematopoietic progenitors to stroma. Blood. 2020;136(23):2607–19.

    Article  PubMed  PubMed Central  Google Scholar 

  122. Thuringer D, Jego G, Berthenet K, Hammann A, Solary E, Garrido C. Gap junction-mediated transfer of miR-145-5p from microvascular endothelial cells to colon cancer cells inhibits angiogenesis. Oncotarget. 2016;7(19):28160–8.

    Article  PubMed  PubMed Central  Google Scholar 

  123. Boengler K, Leybaert L, Ruiz-Meana M, Schulz R. Connexin 43 in mitochondria: what do we really know about its function? Front Physiol. 2022;13: 928934.

    Article  PubMed  PubMed Central  Google Scholar 

  124. Gadicherla AK, Wang N, Bulic M, Agullo-Pascual E, Lissoni A, De Smet M, et al. Mitochondrial Cx43 hemichannels contribute to mitochondrial calcium entry and cell death in the heart. Basic Res Cardiol. 2017;112(3):27.

    Article  PubMed  Google Scholar 

  125. Osswald M, Jung E, Sahm F, Solecki G, Venkataramani V, Blaes J, et al. Brain tumour cells interconnect to a functional and resistant network. Nature. 2015;528(7580):93–8.

    Article  CAS  PubMed  Google Scholar 

  126. Tishchenko A, Azorin DD, Vidal-Brime L, Munoz MJ, Arenas PJ, Pearce C, et al. Cx43 and associated cell signaling pathways regulate tunneling nanotubes in breast cancer cells. Cancers. 2020;12(10):2798.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Li X. Gap junction protein connexin43 and tunneling nanotubes in human trabecular meshwork cells. Int J Physiol Pathophysiol Pharmacol. 2019;11(5):212–9.

    CAS  PubMed  PubMed Central  Google Scholar 

  128. Norris RP. Transfer of mitochondria and endosomes between cells by gap junction internalization. Traffic. 2021;22(6):174–9.

    Article  CAS  PubMed  Google Scholar 

  129. Dixson AC, Dawson TR, Di Vizio D, Weaver AM. Context-specific regulation of extracellular vesicle biogenesis and cargo selection. Nat Rev Mol Cell Biol. 2023;24(7):454–76.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Welsh JA, Holloway JA, Wilkinson JS, Englyst NA. Extracellular vesicle flow cytometry analysis and standardization. Front Cell Dev Biol. 2017;5:78.

    Article  PubMed  PubMed Central  Google Scholar 

  131. Isaac R, Reis FCG, Ying W, Olefsky JM. Exosomes as mediators of intercellular crosstalk in metabolism. Cell Metab. 2021;33(9):1744–62.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. D’Acunzo P, Perez-Gonzalez R, Kim Y, Hargash T, Miller C, Alldred MJ, et al. Mitovesicles are a novel population of extracellular vesicles of mitochondrial origin altered in Down syndrome. Sci Adv. 2021;7(7):eabe5085.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  133. D’Acunzo P, Ungania JM, Kim Y, Barreto BR, DeRosa S, Pawlik M, et al. Cocaine perturbs mitovesicle biology in the brain. J Extracell Vesicles. 2023;12(1): e12301.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Kim Y, Perez-Gonzalez R, Miller C, Kurz M, D’Acunzo P, Goulbourne CN, et al. Sex differentially alters secretion of brain extracellular vesicles during aging: a potential mechanism for maintaining brain homeostasis. Neurochem Res. 2022;47(11):3428–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. She Z, Xie M, Hun M, Abdirahman AS, Li C, Wu F, et al. Immunoregulatory effects of mitochondria transferred by extracellular vesicles. Front Immunol. 2020;11: 628576.

    Article  CAS  PubMed  Google Scholar 

  136. Brestoff JR, Wilen CB, Moley JR, Li Y, Zou W, Malvin NP, et al. Intercellular mitochondria transfer to macrophages regulates white adipose tissue homeostasis and is impaired in obesity. Cell Metab. 2021;33(2):270-82 e8.

    Article  CAS  PubMed  Google Scholar 

  137. Zou X, Kwon SH, Jiang K, Ferguson CM, Puranik AS, Zhu X, et al. Renal scattered tubular-like cells confer protective effects in the stenotic murine kidney mediated by release of extracellular vesicles. Sci Rep. 2018;8(1):1263.

    Article  PubMed  PubMed Central  Google Scholar 

  138. Dave KM, Stolz DB, Venna VR, Quaicoe VA, Maniskas ME, Reynolds MJ, et al. Mitochondria-containing extracellular vesicles (EV) reduce mouse brain infarct sizes and EV/HSP27 protect ischemic brain endothelial cultures. J Control Release. 2023;354:368–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Crewe C, Funcke JB, Li S, Joffin N, Gliniak CM, Ghaben AL, et al. Extracellular vesicle-based interorgan transport of mitochondria from energetically stressed adipocytes. Cell Metab. 2021;33(9):1853-68 e11.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Morrison TJ, Jackson MV, Cunningham EK, Kissenpfennig A, McAuley DF, O’Kane CM, et al. Mesenchymal stromal cells modulate macrophages in clinically relevant lung injury models by extracellular vesicle mitochondrial transfer. Am J Respir Crit Care Med. 2017;196(10):1275–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Hough KP, Trevor JL, Strenkowski JG, Wang Y, Chacko BK, Tousif S, et al. Exosomal transfer of mitochondria from airway myeloid-derived regulatory cells to T cells. Redox Biol. 2018;18:54–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  142. Hough KP, Deshane JS. Exosomes in allergic airway diseases. Curr Allergy Asthma Rep. 2019;19(5):26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  143. Levoux J, Prola A, Lafuste P, Gervais M, Chevallier N, Koumaiha Z, et al. Platelets facilitate the wound-healing capability of mesenchymal stem cells by mitochondrial transfer and metabolic reprogramming. Cell Metab. 2021;33(2):283–99.

    Article  CAS  PubMed  Google Scholar 

  144. Zhao M, Liu S, Wang C, Wang Y, Wan M, Liu F, et al. Mesenchymal stem cell-derived extracellular vesicles attenuate mitochondrial damage and inflammation by stabilizing mitochondrial DNA. ACS Nano. 2021;15(1):1519–38.

    Article  CAS  PubMed  Google Scholar 

  145. Lyamzaev KG, Zinovkin RA, Chernyak BV. Extrusion of mitochondria: garbage clearance or cell-cell communication signals? J Cell Physiol. 2022;237(5):2345–56.

    Article  CAS  PubMed  Google Scholar 

  146. Bagheri HS, Bani F, Tasoglu S, Zarebkohan A, Rahbarghazi R, Sokullu E. Mitochondrial donation in translational medicine; from imagination to reality. J Transl Med. 2020;18(1):367.

    Article  PubMed  PubMed Central  Google Scholar 

  147. Weimann JM, Johansson CB, Trejo A, Blau HM. Stable reprogrammed heterokaryons form spontaneously in Purkinje neurons after bone marrow transplant. Nat Cell Biol. 2003;5(11):959–66.

    Article  CAS  PubMed  Google Scholar 

  148. Noubissi FK, Harkness T, Alexander CM, Ogle BM. Apoptosis-induced cancer cell fusion: a mechanism of breast cancer metastasis. FASEB J. 2015;29(9):4036–45.

    Article  CAS  PubMed  Google Scholar 

  149. Alvarez-Dolado M, Pardal R, Garcia-Verdugo JM, Fike JR, Lee HO, Pfeffer K, et al. Fusion of bone-marrow-derived cells with Purkinje neurons, cardiomyocytes and hepatocytes. Nature. 2003;425(6961):968–73.

    Article  CAS  PubMed  Google Scholar 

  150. Acquistapace A, Bru T, Lesault PF, Figeac F, Coudert AE, le Coz O, et al. Human mesenchymal stem cells reprogram adult cardiomyocytes toward a progenitor-like state through partial cell fusion and mitochondria transfer. Stem cells. 2011;29(5):812–24.

    Article  CAS  PubMed  Google Scholar 

  151. Terada N, Hamazaki T, Oka M, Hoki M, Mastalerz DM, Nakano Y, et al. Bone marrow cells adopt the phenotype of other cells by spontaneous cell fusion. Nature. 2002;416(6880):542–5.

    Article  CAS  PubMed  Google Scholar 

  152. Cusulin C, Monni E, Ahlenius H, Wood J, Brune JC, Lindvall O, et al. Embryonic stem cell-derived neural stem cells fuse with microglia and mature neurons. Stem cells. 2012;30(12):2657–71.

    Article  CAS  PubMed  Google Scholar 

  153. Murray LMA, Krasnodembskaya AD. Concise review: intercellular communication via organelle transfer in the biology and therapeutic applications of stem cells. Stem cells. 2019;37(1):14–25.

    Article  PubMed  Google Scholar 

  154. Lyamzaev KG, Nepryakhina OK, Saprunova VB, Bakeeva LE, Pletjushkina OY, Chernyak BV, et al. Novel mechanism of elimination of malfunctioning mitochondria (mitoptosis): formation of mitoptotic bodies and extrusion of mitochondrial material from the cell. Biochem Biophys Acta. 2008;1777(7–8):817–25.

    CAS  PubMed  Google Scholar 

  155. Nakajima A, Kurihara H, Yagita H, Okumura K, Nakano H. Mitochondrial Extrusion through the cytoplasmic vacuoles during cell death. J Biol Chem. 2008;283(35):24128–35.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Lyamzaev KG, Pletjushkina OY, Saprunova VB, Bakeeva LE, Chernyak BV, Skulachev VP. Selective elimination of mitochondria from living cells induced by inhibitors of bioenergetic functions. Biochem Soc Trans. 2004;32(Pt 6):1070–1.

    Article  CAS  PubMed  Google Scholar 

  157. Lyamzaev KG, Knorre DA, Chernyak BV. Mitoptosis, twenty years after. Biochem Biokhimiia. 2020;85(12):1484–98.

    Article  CAS  Google Scholar 

  158. Nicolas-Avila JA, Lechuga-Vieco AV, Esteban-Martinez L, Sanchez-Diaz M, Diaz-Garcia E, Santiago DJ, et al. A network of macrophages supports mitochondrial homeostasis in the heart. Cell. 2020;183(1):94-109 e23.

    Article  CAS  PubMed  Google Scholar 

  159. Davis CH, Kim KY, Bushong EA, Mills EA, Boassa D, Shih T, et al. Transcellular degradation of axonal mitochondria. Proc Natl Acad Sci USA. 2014;111(26):9633–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Melentijevic I, Toth ML, Arnold ML, Guasp RJ, Harinath G, Nguyen KC, et al. C. elegans neurons jettison protein aggregates and mitochondria under neurotoxic stress. Nature. 2017;542(7641):367–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Unuma K, Aki T, Funakoshi T, Hashimoto K, Uemura K. Extrusion of mitochondrial contents from lipopolysaccharide-stimulated cells: involvement of autophagy. Autophagy. 2015;11(9):1520–36.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Maeda A, Fadeel B. Mitochondria released by cells undergoing TNF-alpha-induced necroptosis act as danger signals. Cell Death Dis. 2014;5(7): e1312.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Boudreau LH, Duchez AC, Cloutier N, Soulet D, Martin N, Bollinger J, et al. Platelets release mitochondria serving as substrate for bactericidal group IIA-secreted phospholipase A2 to promote inflammation. Blood. 2014;124(14):2173–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  164. Ma L, Li Y, Peng J, Wu D, Zhao X, Cui Y, et al. Discovery of the migrasome, an organelle mediating release of cytoplasmic contents during cell migration. Cell Res. 2015;25(1):24–38.

    Article  CAS  PubMed  Google Scholar 

  165. Huang Y, Zucker B, Zhang S, Elias S, Zhu Y, Chen H, et al. Migrasome formation is mediated by assembly of micron-scale tetraspanin macrodomains. Nat Cell Biol. 2019;21(8):991–1002.

    Article  CAS  PubMed  Google Scholar 

  166. Jiao H, Jiang D, Hu X, Du W, Ji L, Yang Y, et al. Mitocytosis, a migrasome-mediated mitochondrial quality-control process. Cell. 2021;184(11):2896–910.

    Article  CAS  PubMed  Google Scholar 

  167. Yu S, Yu L. Migrasome biogenesis and functions. FEBS J. 2022;289(22):7246–54.

    Article  CAS  PubMed  Google Scholar 

  168. Savukinas UB, Enes SR, Sjoland AA, Westergren-Thorsson G. Concise review: the bystander effect: mesenchymal stem cell-mediated lung repair. Stem cells. 2016;34(6):1437–44.

    Article  PubMed  Google Scholar 

  169. Rustom A, Saffrich R, Markovic I, Walther P, Gerdes HH. Nanotubular highways for intercellular organelle transport. Science. 2004;303(5660):1007–10.

    Article  CAS  PubMed  Google Scholar 

  170. Jiang D, Chen FX, Zhou H, Lu YY, Tan H, Yu SJ, et al. Bioenergetic crosstalk between mesenchymal stem cells and various ocular cells through the intercellular trafficking of mitochondria. Theranostics. 2020;10(16):7260–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  171. Han D, Zheng X, Wang X, Jin T, Cui L, Chen Z. Mesenchymal stem/stromal cell-mediated mitochondrial transfer and the therapeutic potential in treatment of neurological diseases. Stem cells international. 2020;2020:8838046.

    Article  PubMed  PubMed Central  Google Scholar 

  172. Regmi S, Liu DD, Shen M, Kevadiya BD, Ganguly A, Primavera R, et al. Mesenchymal stromal cells for the treatment of Alzheimer’s disease: Strategies and limitations. Front Mol Neurosci. 2022;15:1011225.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Ben-Shachar D, Laifenfeld D. Mitochondria, synaptic plasticity, and schizophrenia. Int Rev Neurobiol. 2004;59:273–96.

    Article  CAS  PubMed  Google Scholar 

  174. Sun T, Ma QH. Repairing neural injuries using human umbilical cord blood. Mol Neurobiol. 2013;47(3):938–45.

    Article  CAS  PubMed  Google Scholar 

  175. Jurkunas UV. Fuchs endothelial corneal dystrophy through the prism of oxidative stress. Cornea. 2018;37(Suppl 1):S50–4.

    Article  PubMed  Google Scholar 

  176. Kumar B, Kowluru A, Kowluru RA. Lipotoxicity augments glucotoxicity-induced mitochondrial damage in the development of diabetic retinopathy. Invest Ophthalmol Vis Sci. 2015;56(5):2985–92.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  177. Zhong Q, Kowluru RA. Diabetic retinopathy and damage to mitochondrial structure and transport machinery. Invest Ophthalmol Vis Sci. 2011;52(12):8739–46.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Jiang D, Xiong G, Feng H, Zhang Z, Chen P, Yan B, et al. Donation of mitochondria by iPSC-derived mesenchymal stem cells protects retinal ganglion cells against mitochondrial complex I defect-induced degeneration. Theranostics. 2019;9(8):2395–410.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Subramaniam MD, Chirayath RB, Iyer M, Nair AP, Vellingiri B. Mesenchymal stem cells (MSCs) in Leber’s hereditary optic neuropathy (LHON): a potential therapeutic approach for future. Int Ophthalmol. 2022;42(9):2949–64.

    Article  PubMed  Google Scholar 

  180. Crestini A, Santilli F, Martellucci S, Carbone E, Sorice M, Piscopo P, et al. Prions and neurodegenerative diseases: a focus on Alzheimer’s disease. Journal of Alzheimer’s disease : JAD. 2022;85(2):503–18.

    Article  CAS  PubMed  Google Scholar 

  181. Lee NK, Park SE, Kwon SJ, Shim S, Byeon Y, Kim JH, et al. Agouti related peptide secreted via human mesenchymal stem cells upregulates proteasome activity in an Alzheimer’s disease model. Sci Rep. 2017;7:39340.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Mendonca CF, Kuras M, Nogueira FCS, Pla I, Hortobagyi T, Csiba L, et al. Proteomic signatures of brain regions affected by tau pathology in early and late stages of Alzheimer’s disease. Neurobiol Dis. 2019;130: 104509.

    Article  CAS  PubMed  Google Scholar 

  183. Neves AF, Camargo C, Premer C, Hare JM, Baumel BS, Pinto M. Intravenous administration of mesenchymal stem cells reduces Tau phosphorylation and inflammation in the 3xTg-AD mouse model of Alzheimer’s disease. Exp Neurol. 2021;341: 113706.

    Article  CAS  PubMed  Google Scholar 

  184. Manczak M, Calkins MJ, Reddy PH. 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(13):2495–509.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Manczak M, Reddy PH. Abnormal interaction between the mitochondrial fission protein Drp1 and hyperphosphorylated tau in Alzheimer’s disease neurons: implications for mitochondrial dysfunction and neuronal damage. Hum Mol Genet. 2012;21(11):2538–47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Lin MT, Beal MF. Mitochondrial dysfunction and oxidative stress in neurodegenerative diseases. Nature. 2006;443(7113):787–95.

    Article  CAS  PubMed  Google Scholar 

  187. Kim JY, Kim DH, Kim JH, Lee D, Jeon HB, Kwon SJ, et al. Soluble intracellular adhesion molecule-1 secreted by human umbilical cord blood-derived mesenchymal stem cell reduces amyloid-beta plaques. Cell Death Differ. 2012;19(4):680–91.

    Article  CAS  PubMed  Google Scholar 

  188. Sha S, Shen X, Cao Y, Qu L. Mesenchymal stem cells-derived extracellular vesicles ameliorate Alzheimer’s disease in rat models via the microRNA-29c-3p/BACE1 axis and the Wnt/beta-catenin pathway. Aging. 2021;13(11):15285–306.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Yokokawa K, Iwahara N, Hisahara S, Emoto MC, Saito T, Suzuki H, et al. Transplantation of mesenchymal stem cells improves amyloid-beta pathology by modifying microglial function and suppressing oxidative stress. J Alzheimer’s Dis JAD. 2019;72(3):867–84.

    Article  CAS  PubMed  Google Scholar 

  190. Cui Y, Ma S, Zhang C, Cao W, Liu M, Li D, et al. Human umbilical cord mesenchymal stem cells transplantation improves cognitive function in Alzheimer’s disease mice by decreasing oxidative stress and promoting hippocampal neurogenesis. Behav Brain Res. 2017;320:291–301.

    Article  CAS  PubMed  Google Scholar 

  191. Zhang Z, Sheng H, Liao L, Xu C, Zhang A, Yang Y, et al. Mesenchymal stem cell-conditioned medium improves mitochondrial dysfunction and suppresses apoptosis in okadaic acid-treated SH-SY5Y cells by extracellular vesicle mitochondrial transfer. J Alzheimer’s Dis JAD. 2020;78(3):1161–76.

    Article  CAS  PubMed  Google Scholar 

  192. Alves AA, Andrade LM, Barbosa AF, Bediaga I, Cernicchiaro G, Guerrer G, et al. The LHCb Detector at the LHC. J Instrum. 2008;3:S08005.

    Google Scholar 

  193. Oyinbo CA. Secondary injury mechanisms in traumatic spinal cord injury: a nugget of this multiply cascade. Acta Neurobiol Exp. 2011;71(2):281–99.

    Article  Google Scholar 

  194. Turrens JF. Mitochondrial formation of reactive oxygen species. J Physiol. 2003;552(Pt 2):335–44.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Scholpa NE, Schnellmann RG. Mitochondrial-based therapeutics for the treatment of spinal cord injury: mitochondrial biogenesis as a potential pharmacological target. J Pharmacol Exp Ther. 2017;363(3):303–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Shende P, Subedi M. Pathophysiology, mechanisms and applications of mesenchymal stem cells for the treatment of spinal cord injury. Biomed Pharmacother Biomed Pharmacother. 2017;91:693–706.

    Article  CAS  PubMed  Google Scholar 

  197. Lin L, Lin H, Bai S, Zheng L, Zhang X. Bone marrow mesenchymal stem cells (BMSCs) improved functional recovery of spinal cord injury partly by promoting axonal regeneration. Neurochem Int. 2018;115:80–4.

    Article  CAS  PubMed  Google Scholar 

  198. Yang EZ, Zhang GW, Xu JG, Chen S, Wang H, Cao LL, et al. Multichannel polymer scaffold seeded with activated Schwann cells and bone mesenchymal stem cells improves axonal regeneration and functional recovery after rat spinal cord injury. Acta Pharmacol Sin. 2017;38(5):623–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  199. Poudineh M, Wang Z, Labib M, Ahmadi M, Zhang L, Das J, et al. Three-dimensional nanostructured architectures enable efficient neural differentiation of mesenchymal stem cells via mechanotransduction. Nano Lett. 2018;18(11):7188–93.

    Article  CAS  PubMed  Google Scholar 

  200. Sykova E, Jendelova P, Urdzikova L, Lesny P, Hejcl A. Bone marrow stem cells and polymer hydrogels–two strategies for spinal cord injury repair. Cell Mol Neurobiol. 2006;26(7–8):1113–29.

    CAS  PubMed  Google Scholar 

  201. Deng WS, Ma K, Liang B, Liu XY, Xu HY, Zhang J, et al. Collagen scaffold combined with human umbilical cord-mesenchymal stem cells transplantation for acute complete spinal cord injury. Neural Regen Res. 2020;15(9):1686–700.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Kalogeris T, Baines CP, Krenz M, Korthuis RJ. Ischemia/reperfusion. Compr Physiol. 2016;7(1):113–70.

    Article  PubMed  PubMed Central  Google Scholar 

  203. Anzell AR, Maizy R, Przyklenk K, Sanderson TH. Mitochondrial quality control and disease: insights into ischemia-reperfusion injury. Mol Neurobiol. 2018;55(3):2547–64.

    Article  CAS  PubMed  Google Scholar 

  204. Liu K, Ji K, Guo L, Wu W, Lu H, Shan P, et al. Mesenchymal stem cells rescue injured endothelial cells in an in vitro ischemia-reperfusion model via tunneling nanotube like structure-mediated mitochondrial transfer. Microvasc Res. 2014;92:10–8.

    Article  CAS  PubMed  Google Scholar 

  205. Murphy E, Steenbergen C. Mechanisms underlying acute protection from cardiac ischemia-reperfusion injury. Physiol Rev. 2008;88(2):581–609.

    Article  CAS  PubMed  Google Scholar 

  206. Halestrap AP. Calcium, mitochondria and reperfusion injury: a pore way to die. Biochem Soc Trans. 2006;34(Pt 2):232–7.

    Article  CAS  PubMed  Google Scholar 

  207. Luongo TS, Lambert JP, Gross P, Nwokedi M, Lombardi AA, Shanmughapriya S, et al. The mitochondrial Na(+)/Ca(2+) exchanger is essential for Ca(2+) homeostasis and viability. Nature. 2017;545(7652):93–7.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  208. Soundara Rajan T, Gugliandolo A, Bramanti P, Mazzon E. Tunneling nanotubes-mediated protection of mesenchymal stem cells: an update from preclinical studies. Int J Mol Sci. 2020;21(10):3481.

    Article  PubMed  PubMed Central  Google Scholar 

  209. Yamada Y, Ito M, Arai M, Hibino M, Tsujioka T, Harashima H. Challenges in promoting mitochondrial transplantation therapy. Int J Mol Sci. 2020;21(17):6365.

    Article  PubMed  PubMed Central  Google Scholar 

  210. McCully JD, Cowan DB, Pacak CA, Toumpoulis IK, Dayalan H, Levitsky S. Injection of isolated mitochondria during early reperfusion for cardioprotection. Am J Physiol Heart Circ Physiol. 2009;296(1):H94–105.

    Article  CAS  PubMed  Google Scholar 

  211. Masuzawa A, Black KM, Pacak CA, Ericsson M, Barnett RJ, Drumm C, et al. Transplantation of autologously derived mitochondria protects the heart from ischemia-reperfusion injury. Am J Physiol Heart Circ Physiol. 2013;304(7):H966–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. Kaza AK, Wamala I, Friehs I, Kuebler JD, Rathod RH, Berra I, et al. Myocardial rescue with autologous mitochondrial transplantation in a porcine model of ischemia/reperfusion. J Thorac Cardiovasc Surg. 2017;153(4):934–43.

    Article  PubMed  Google Scholar 

  213. Guariento A, Blitzer D, Doulamis I, Shin B, Moskowitzova K, Orfany A, et al. Preischemic autologous mitochondrial transplantation by intracoronary injection for myocardial protection. J Thorac Cardiovasc Surg. 2020;160(2):e15–29.

    Article  PubMed  Google Scholar 

  214. Blitzer D, Guariento A, Doulamis IP, Shin B, Moskowitzova K, Barbieri GR, et al. Delayed transplantation of autologous mitochondria for cardioprotection in a porcine model. Ann Thorac Surg. 2020;109(3):711–9.

    Article  PubMed  Google Scholar 

  215. Emani SM, Piekarski BL, Harrild D, Del Nido PJ, McCully JD. Autologous mitochondrial transplantation for dysfunction after ischemia-reperfusion injury. J Thorac Cardiovasc Surg. 2017;154(1):286–9.

    Article  PubMed  Google Scholar 

  216. O’Brien CG, Ozen MO, Ikeda G, Vaskova E, Jung JH, Bayardo N, et al. Mitochondria-rich extracellular vesicles rescue patient-specific cardiomyocytes from doxorubicin injury: insights into the SENECA trial. JACC CardioOncol. 2021;3(3):428–40.

    Article  PubMed  PubMed Central  Google Scholar 

  217. Maremanda KP, Sundar IK, Rahman I. Protective role of mesenchymal stem cells and mesenchymal stem cell-derived exosomes in cigarette smoke-induced mitochondrial dysfunction in mice. Toxicol Appl Pharmacol. 2019;385: 114788.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  218. Kubat GB, Ozler M, Ulger O, Ekinci O, Atalay O, Celik E, et al. The effects of mesenchymal stem cell mitochondrial transplantation on doxorubicin-mediated nephrotoxicity in rats. J Biochem Mol Toxicol. 2021;35(1): e22612.

    Article  CAS  PubMed  Google Scholar 

  219. Lu T, Zhang J, Cai J, Xiao J, Sui X, Yuan X, et al. Extracellular vesicles derived from mesenchymal stromal cells as nanotherapeutics for liver ischaemia-reperfusion injury by transferring mitochondria to modulate the formation of neutrophil extracellular traps. Biomaterials. 2022;284: 121486.

    Article  CAS  PubMed  Google Scholar 

  220. Bi Y, Guo X, Zhang M, Zhu K, Shi C, Fan B, et al. Bone marrow derived-mesenchymal stem cell improves diabetes-associated fatty liver via mitochondria transformation in mice. Stem Cell Res Ther. 2021;12(1):602.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Funding

No funding.

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization, I.R., V.M. and S.D.M.; writing—original draft preparation, S.P, I.R., S.D.M. and V.M.; writing—review and editing, S.P., I.R., V.M. and S.D.M.; supervision, I.R., V.M. and S.D.M. All authors have read and agreed to the published version of the manuscript.

Corresponding authors

Correspondence to Vincenzo Mattei or Simona Delle Monache.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Iorio, R., Petricca, S., Mattei, V. et al. Horizontal mitochondrial transfer as a novel bioenergetic tool for mesenchymal stromal/stem cells: molecular mechanisms and therapeutic potential in a variety of diseases. J Transl Med 22, 491 (2024). https://doi.org/10.1186/s12967-024-05047-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12967-024-05047-4

Keywords