Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2023 Dec 1.
Published in final edited form as: Trends Cancer. 2022 Jul 29;8(12):1002–1018. doi: 10.1016/j.trecan.2022.07.004

Roles of mitochondrial genetics in cancer metastasis

Danny R Welch 1,2,3,4,5,*, Christian Foster 1, Isidore Rigoutsos 6
PMCID: PMC9884503  NIHMSID: NIHMS1867429  PMID: 35915015

Abstract

The contributions of mitochondria to cancer have been recognized for decades. However, the focus on the metabolic role of mitochondria and the diminutive size of the mitochondrial genome compared to the nuclear genome have hindered discovery of the roles of mitochondrial genetics in cancer. This review summarizes recent data demonstrating the contributions of mitochondrial DNA (mtDNA) copy-number variants (CNVs), somatic mutations, and germline polymorphisms to cancer initiation, progression, and metastasis. The goal is to summarize accumulating data to establish a framework for exploring the contributions of mtDNA to neoplasia and metastasis.

Mitochondrial contributions to disease

Cancers are polygenic diseases with genes driving oncogenesis, suppressing tumorigenicity, or modifying disease severity. Most research has focused on nuclear genome, mostly because of its relative size and greater level of characterization. However, eukaryotic cells also possess a genome within the mitochondrial organelle. Mitochondria are referred to as the ‘powerhouse of the cell’ because they are central to many metabolic and catabolic processes. Accumulating data demonstrate that mitochondria can no longer be seen simply as energy plants because they contribute to complex phenotypes in previously underappreciated ways [13].

To understand how mitochondrial genetics contribute to cancer, it is first important to define the origins and evolution of the mitochondrial organelle (Box 1) and the unique characteristics of the mitochondrial genome (Box 2). It is beyond the scope of this review to do so comprehensively, but a few key principles will be crucial in dissecting its contributions in cancer. (i) mtDNA is maternally inherited, and several different SNPs have been selected through evolution and can be used to distinguish ancestral groups. (ii) Copies of the circular mtDNA genome vary greatly between cells and tissues. (iii) When cells divide, mtDNA is not equally divided between progenitor cells. (iv) Mitochondria are dynamic organelles, changing shape and position according to physiologic needs (Box 3). When assessing mitochondrial contributions to disease, each of these characteristics must be considered. Consequently, SNPs and somatic mutations in mtDNA have been associated with age-related pathologies, for example neurodegenerative disorders and cancers [46]. Together these observations lead to the question: do mtDNA mutations or adaptations associated with human ancestry impact on disease susceptibility, progression, or severity? The goal of this brief review is to summarize recent data and establish a framework for dissecting the contributions of mitochondria (mtDNA SNPs and/or mutations) in neoplasia and metastasis (see Glossary).

Box 1. Mitochondrial evolution.

In 1905 the Russian botanist Konstantin Mereschkowski first postulated that the chloroplast was the result of a symbiotic relationship in plants [7]. In 1967, Lynn Sagan (née Margulis) proposed that the related organelle, the mitochondrion, was similarly symbiotic [8]. These concepts have been expanded to propose that mitochondria arose from incorporation of an Alphaproteobacterium into a host cell over a billion years ago [9]. From which, subsequent coevolution led to added functions to meet cellular needs, including intercellular communication [10,11], epigenetic tagging [1214], innate immunity [15], regulation of apoptosis [16], and involvement in disease.

Current theories surrounding SNPs comprising human haplotypes support the existence of a mitochondrial ‘Eve’ from which matrilineal selection of variants based upon energy demands and thermoregulation occurred when humans migrated from Africa through Europe to Asia, the Americas, and Australasia [9,1723]. As a result, SNPs which define haplogroups can be used as a criterion by which to recognize human population groups (i.e., ancestry or colloquially ‘race’).

Box 2. Mitochondrial DNA.

Mammalian mitochondrial genomes consist of ~16.5 kb of circular DNA that encodes 13 protein subunits of the electron transport chain (ETC), two rRNAs (16S and 12S), and 22 tRNAs. Replication of the mtDNA starts in the displacement (D)-loop, which also contains transcriptional control regions for mtDNA. Some bases within the D-loop region are conserved whereas others are highly variable. The latter have proved to be useful for studying vertebrate evolution. Mammalian mitochondrial genomes are similar but not identical. Table I compares human and murine mtDNA as an example. The strands of the mtDNA duplex can be distinguished based upon base composition. The heavy (H) strand contains more guanine and is the template from which 12 of the 13 mitochondrial protein mRNAs are transcribed, whereas the light (L) strand contains more cytosines and encodes only one protein [24].

Within cells, there can be tens to thousands of mtDNA copies [1,25,26]. Over time, the originally engulfed bacterium has ceded much of its genome to nDNA. Nuclear genes contribute >90% of the molecules responsible for mitochondrial structure, function, bioenergetics, replication, and repair [27,28]. A consequence is protection of crucial molecules from the 10- to 100-fold higher mutation rates found in mtDNA compared to nDNA [19]. Recently, it was reported that seven mitochondrial tRNAs have identical copies embedded in the human nuclear genome [29], presumably as a protection against mutations. Evidently this phenomenon is also present in animals – at least in marsupials [30], and plants (I.R., unpublished). The primary contributor to mtDNA mutations is the high level of reactive oxygen species (ROS) generated during electron transport accompanied by lower-fidelity DNA replication mediated by polymerase γ (Pol γ) [31,32] and Pol τ under conditions of oxidative stress [33]. Although Pol γ has high fidelity, an increased mutation rate is observed because of an elevated rate of replication rate, limited repair mechanisms, high ROS, and lack of protective histones [3437]. A vast number of mtDNA mutations arise from DNA replication errors with heavy- and light-strand biases [38]. mtDNA includes regions with abundant nucleotide repeats which have been attributed to increased mtDNA mutations because of DNA polymerase slippage leading to the introduction of run-length variation at these sites [38,39]. Increased ROS levels have also been demonstrated to correlate with increased mtDNA mutation burden and tumor progression [40].

In addition, it was recently shown that mitochondrially encoded tRNAs correlate positively with nuclear mRNAs whose exons and introns contain repetitive elements [93]. This finding links repetitive elements to cancer via mtDNA. Notably, recent work has shown that mt-tRNA-derived fragments are associated with the nuclear localization, membrane localization, and secretion of specific proteins [93,94]. The details of these associations and the identity of the involved proteins differ by cancer type and are the subject of current investigation.

Box 3. Mitochondrial dynamics.

Some cells require longer mitochondrial filaments to fulfill cellular metabolic needs, whereas others have less physical space for organelles to function and are more uniform and spherical to fit into smaller spaces (e.g., hepatocytes). Although mitochondria within specific tissues have characteristic phenotypes, morphology is not static. The balance of fission and fusion changes in response to growth, mitochondrial damage, and energy levels [41,42]. These processes are crucial for the survival of all types of cells, even senescent cells [43]. For example, nonproliferating neurons require a proper balance of fission and fusion for normal physiological function because, without it, disease states such as Charcot–Marie–Tooth disease [44] and Leber’s hereditary optic neuropathy (LHON) can occur [45,46]. Embryogenesis cannot proceed and is often fatal when mechanisms connected to fission/fusion are dysfunctional. Fission is essential for growing and dividing cells to ensure organelle distribution between progeny, whereas fusion is key to maintaining mitochondrial health. Generally, when energy needs are increased, fusion promotes OXPHOS when protein synthesis is inhibited by starvation. Mitochondria fuse, mixing content in an apparent effort to correct damage or complement needs of malfunctioning mitochondria. When fission/fusion mechanisms are broken or hijacked because of acquired mutations, pathologies emerge, including cancers [47,48]. Recent data reveal that the subcellular localization of mitochondria also varies widely during cellular processes including migration and invasion that are most pertinent to this review [4953]. It is worth interjecting that both Charcot–Marie–Tooth and LHON have also been linked to tRNA synthesis and functions as well as to mitochondrial dynamics [5457].

Mitochondrial diseases

Mitochondrial diseases manifest largely in tissues which rely heavily on metabolism (e.g., skeletal and cardiac muscle) or in tissues that rely on glucose as their major energy source (e.g., brain and neurons) [710]. The latter result in LHON (Leber’s hereditary optic neuropathy), MELAS (mitochondrial encephalopathy, lactic acidosis, and stroke-like episodes), MERFF (myoclonic epilepsy with ragged red fibers), and NARP (neuropathy, ataxia, and retinitis pigmentosa) syndromes [3,11,12]. Mitochondria play many roles in other common diseases such as Alzheimer’s, Parkinson’s, glaucoma, multiple sclerosis, amyotrophic lateral sclerosis, cardiovascular disease, osteoarthritis, bipolar disorder, and type 2 diabetes [2,3], but those roles are typically secondary in nature and, at present, remain relatively ill-defined. Neoplastic cells are often metabolically unbalanced, eventually accumulating changes that snowball into the late-stage phenotypes observed clinically. Cancers are increasingly included among diseases in which mitochondrial mutations are observed [3,1319], but whether those mutations are the cause or consequence (or both) is still under investigation.

The severity of most mitochondrial diseases is related to a ‘biochemical threshold’ which is when the number of mutant copies surpasses a given ratio within a cell or group of cells. The threshold can be widely distributed (i.e., affecting multiple tissues) or in specific tissues. This heterogeneity is, in part, why mitochondrial diseases are challenging to diagnose and understand. In some cells the normal, unmutated mtDNA can protect cells from mutated variants by rescue mechanisms such as fission/fusion [7,11,20]. Shifts in mtDNA content can alter metabolism and nuclear epigenetic marks [21,22]. It is also possible that mtDNA epigenetic modifications are also affected by mtDNA copy number, but to the best of our knowledge this has not been measured.

At birth, mtDNA is generally identical by sequence in all cells (homoplasmy), although copy numbers can vary by tissue. With aging, mtDNA can accumulate mutations such that cells contain mixtures of mtDNA (heteroplasmy). Because mtDNA is not distributed equally among progeny, heteroplasmy contributes to cellular evolution [11]. Further, heteroplasmy impacts on disease severity and disease subtype [20,23]. Interestingly, heteroplasmy levels fluctuate depending on tissue type and cellular energy needs, but revert toward homoplasmy that provides a selective advantage over another mutation. As a result, it is crucial to consider both sequence and copy number when assessing the contributions of mtDNA in cancer. From a technical perspective, the increased depth of genome sequencing has resulted in increased detection of heteroplasmy [4,5] which in turn complicates definitive attribution of driver versus ancillary mutations compared to the mostly diploid composition of nucleus-encoded genes. As a rule, persistent heteroplasmy contributes to phenotype instability and reduced cellular fitness [6]. Thus, there is selection pressure to reduce heteroplasmy even within cancer cells [24,25], which suggests that homoplasmy may be important for maintenance of mitochondrial function. Sercel and colleagues recently evaluated homo- and heteroplasmy in iPSCs and showed that, for retention of pluripotency, homoplasmy was a crucial determinant [26]. Whether the same patterns are observed in cancer stem cells has not yet been measured to the best of our knowledge.

Metabolism and cancer

The primary functions of mitochondria relate to energy homeostasis. Electron transport achieves this through a series of redox reactions that shuffle electrons and pump protons across the inner membrane, driving ATP synthesis via proton motive forces. The ‘tighter’ the complex, the more ATP is produced, the more reactive oxygen species (ROS) are produced, and less heat is generated. The loosely packed electron transport chain (ETC) (i.e., uncoupled) exhibits less efficient ATP production, reduced ROS levels, and greater heat emission. Uncoupling respiration from ATP synthesis was thought to be one the most crucial adaptations of Homo spp. living in colder climates. Several proteins responsible for uncoupling (uncoupling proteins, UCPs) have been implicated in cancer but also vary by tissue and uncoupling is influenced by the microenvironment.

In the 1920s Warburg reported a phenomenon referred to as ‘aerobic glycolysis’ in which neoplastic cells shift glucose conversion to ATP via glycolysis rather than via oxidative phosphorylation (OXPHOS) despite adequate oxygenation [2729]. Subsequently, these observations have been replicated, but the hypothesis that mitochondrial dysfunction is a driver of cancer needed to be modified because most cancer cells maintain mitochondrial respiration [30,31]. Nevertheless, Warburg’s observations ushered in the concept that mitochondria are pivotal to cancer biology because of their regulation of metabolism. In addition, it is now appreciated that gain-of-function mutations in isodehydrogenase 1/2 (IDH1/2) can drive cancer [32]. Despite the discovery that some nuclear (n)DNA-encoded mitochondrial proteins regulate tumorigenesis [33,34], mtDNA SNPs or mutations are generally not considered to be driver mutations [3537]. A rare benign class of tumors, termed oncocytomas, are an exception. They are characterized by abnormal accumulations of dysfunctional mitochondria [38], or accumulations of normal mitochondria that have increased respiration due to loss of a tumor-suppressor gene FLCN in the nucleus, both of which can cause malignancy.

mtDNA mutations in cancer and metastasis

It is important to emphasize that mtDNA encodes quantitative trait loci (QTLs) that work with nucleus-encoded genes to regulate many complex diseases [3941]. Measurable phenotypic changes are influenced by the combined effects of SNPs, mutations, and environmental factors. It is therefore unlikely that mitochondrial polymorphisms will provide a complete explanation of differences in disease progression. These SNPs will, however, influence disease development and progression by gene–gene interactions and altering responses to the tumor microenvironment. This is a crucial point because mitochondrial SNPs would otherwise be predicted to exhibit strong maternal inheritance because mtDNA is maternally inherited. Instead, crosstalk between the nucleus and mitochondria conveys signals from the various microenvironments in which disseminated cells find themselves during the metastatic cascade. Depending upon the signals and the response of each cell, the efficiency of metastasis will be altered. Likewise, it must be emphasized that not all SNPs in mtDNA will be QTLs for a given phenotype.

The designation of mtDNA mutations as drivers has been difficult because of the limitations of experimental models and technology. Intriguingly, in a study evaluating breast cancers, mtDNA mutational burden was not correlated with survival [42]. Nevertheless, The Cancer Genome Atlas (TCGA) datasets identified some fascinating correlations [24,25] that will be elaborated later together with more recent findings. Definitive cause–effect relationships, however, are difficult to ascribe owing to the existence of multiple mtDNA copies per cell and the technical challenges associated with manipulating all copies of mtDNA [11,43,44]. In addition, most next-generation sequencing (NGS) studies filter out mtDNA sequences, possibly because the small mitochondrial genome was suspected to be inconsequential. Those raw data still exist and are expected to be a goldmine of information as the relevance of the mitochondrial genome, not only of metabolism, in cancer is appreciated more fully.

Mitochondrial CNVs correlate with cancer and metastasis

Recent evidence shows that CNVs exist for mtDNA in multiple cancers [2,13,37,45] (Figure 1, Key Figure). Ovarian cancers tended to have the most (>600 copies) whereas myeloid cancers had the least (~90 copies). Comparing cancer to matched non-cancerous tissues, increased mtDNA was found in patients with chronic lymphocytic leukemia, lung squamous cell carcinoma, or pancreatic adenocarcinoma, whereas the pattern was reversed in kidney clear cell carcinoma, hepatocellular carcinoma, and myeloproliferative neoplasms. CNVs correlated positively with age in prostate, colorectal, and skin cancer [37]. Interestingly, restoration of KISS1 metastasis-suppressor expression corresponded to an increase in the mitochondrial biogenesis regulator, PGC1α, with a corresponding increase in mitochondrial mass in the tumor cells [46]. Collectively, these data show that mitochondrial biogenesis regulation can result in apparently discordant results that differentially regulate oncogenesis and/or metastasis. Although the biogenesis pathways are clearly changed, definitive cause–effect relationships have not been established.

Figure 1. Key figure. Mitochondrial genetics contribute to cancer.

Figure 1. Key figure

Mitochondrial (mt)DNA copy-number variants (CNVs) or germline SNPs are associated with the development and progression to metastasis of multiple cancer histotypes. In addition, somatic mutations emerge and are associated with progression of most cancers, but their roles as drivers are not established. Note: it is possible for SNPs and CNVs to coexist, but those data have not yet been widely studied nor reported. Abbreviation: osteoSa, osteosarcoma.

Mitochondrial transfer between cells

In recent years the mitochondrial content of tumor and neighboring host cells has been shown to change through exchange of mtDNA or whole mitochondria between cells via tunneling nanotubes (TNTs) [4754]. Transfer has been associated with proliferation, aggressiveness, and glycolysis, as well as with pentose phosphate and lipid metabolism, in a variety of neoplastic cells [4854]. These changes often occur concomitantly with chemoresistance [48,50,55]. It is currently only speculation what triggers TNTs, but cellular attempts to repair mtDNA damage is among the possible reasons. In a recent study, transfer of normal mitochondria into cisplatin-resistant MDA-MB-231 breast carcinoma cells restored chemosensitivity [56], leading the authors to speculate that mitochondrial transfer should be considered as a therapeutic option. Although an intriguing possibility, technical hurdles related to efficiency of transfer and instability associated with introduced heteroplasmy will need to be more thoroughly considered beforehand. Relatedly, recent data suggest that elevated levels of uncoupled protein 2 (UCP2) may contribute to chemoresistance [57] but this requires further investigation.

Germline mtDNA polymorphisms may explain cancer disparities

In 2020 Yuan et al. [37] performed a meta-analysis aggregating whole-genome sequencing data from 2658 cancers across 38 tumor types to identify germline and somatically acquired mutations coupled with assessment of transcriptional activities of mitochondrial genes. Collectively, they confirmed many of the patterns described earlier. They found that MT-ND5 was the most frequently mutated ETC gene in most cancers, whereas MT-ND4 was most frequently mutated in prostate and lung cancers, and MT-COX1 was most frequently mutated in breast, cervical, and bladder cancers. Most mutations in tumor types were similar, with C:G>T:A in >50% of cases. mtDNA somatic mutations were acquired at an early age and shifted toward homoplasmy throughout life in the cellular lineage of the neoplastic cells. Whether the shift to homoplasmy was due to asymmetric segregations during cell division or by selection of the mutations remains unknown. Interestingly, in the subset of kidney and thyroid cancers with no identifiable oncogenic driver, they found non-silent mtDNA mutations, suggesting a potential contribution of these mutations in the absence of nuclear drivers. They also observed negative selection for truncating mutations of mtDNA-encoded proteins, suggesting the importance of an intact ETC for most cancer cells. Exceptions were found in kidney, colorectal, and thyroid cancers, possibly implicating inactivation of mitochondrial genes in tumorigenesis for some cell types.

Although the aforementioned analyses are informative, a key conclusion is that mitochondrial contributions to cancer vary by tissue of origin, cancer subtype, age, microenvironmental stress, and other so far unidentified variables. Therefore, it is instructive to examine several of the individual studies that contributed to the analysis by Yuan et al. as well as some newer data obtained since their publication.

NADH-ubiquinone oxidoreductase (complex I) is the largest complex of the ETC. It is essential for biosynthesis and redox control during proliferation, resistance to cell death, and metastasis [5861]. Several inhibitors of complex I act as selective anticancer agents. Paradoxically, other studies show that complex I inhibition shows protumor effects [59]. Similarly, mutations in MT-ND1, that encodes a component of complex I, can enhance non-small cell lung cancer (NSCLC) metastasis [62]. Mito-mice carrying the G13997A mutation, which induces complex I, exhibited high lactate production but no observable phenotypes [63], suggesting that additional genetic or epigenetic cofactors are necessary to influence tumorigenicity and metastasis, consistent with our assertions that mtDNA SNPs are QTLs.

Oncogenes contribute to altered metabolism [6466]. For example, mutant p53 with the P72R gain-of-function mutation regulates PGC1α owing to changes in binding and leads to poorer prognosis [67]. In addition, c-Myc can cause mtDNA fragmentation [68]. Knowing this, it is reasonable to ask: do oncogenes alter tumorigenesis in the context of mtDNA mutations or SNPs? Do the combinatorial effects of mtDNA QTLs and nDNA QTLs explain differential susceptibilities to cancer and/or metastasis? mtDNA mutations do not occur with equal frequency or location in all cancer types – prostate, liver, stomach, and colorectal cancers have the highest, whereas heme malignancies generally have the fewest [25]. We previously posed several questions that arose from these observations [69]. Are there selective pressures for specific mitochondrial mutations? Are there tissue-specific variations for mtDNA mutations in oncogenesis or metastasis? Are there ‘hotspots’ for mutation in cancer (progression)? Are there germline differences in mtDNA (i.e., SNPs) which could (partially) explain racial disparities in cancer and/or metastasis development? (see Outstanding questions). Tables 1 and 2 show changes in relative risk associated with mtDNA haplotypes.

Outstanding questions.

What selective pressures drive the acquisition of, or selection for, mtDNA mutations?

Are there ‘hotspots’ in mtDNA that are mutated in most cancers, in select cancers, or in embryologically related cancers?

Are somatic mtDNA mutations drivers, passengers, or hitchhikers?

Are there germline differences in mtDNA (i.e., SNPs) that predispose people to cancer and/or metastasis, or that can predict responses to therapy?

Do mutations or SNPs in mtDNA regulate subcellular distribution or the transfer of mitochondria or mtDNA between cells?

Do mutations in nDNA disrupt the production of mtRNA?

What are the signals to and from the mitochondria that mediate the phenotypic changes in cancer?

What is the molecular basis for interactions between the mitochondrial genome and the nuclear genome?

Table 1.

mtDNA mutations and polymorphisms in cancera

Cancer type Gene Posion Cancer type Gene Posion
Bladder MT-ND2 T4823C NSCLC MT-ND5 C12813A
Breast MT-CYB A15836C NSCLC MT-ND4 C11409T
Breast MT-CYB T14819insTTCTATA NSCLC MT-ND3 T10363C
Breast MT-ND5 G13333A NSCLC MT-ND1 T3394C
Breast MT-ATP6 T8821C NSCLC MT-ND1 C3497T
Breast MT-CO3 A9664G NSCLC MT-ND1 C3689G
Colorectal MT-ND6 T14470C Ovary D-loop C16296T
Colorectal MT-ND1 C3497T Ovary D-loop C16294T
Colorectal MT-ND1 T3394C Ovary D-loop C16261T
Colorectal NMT-D5 G12630A Ovary D-loop A16163G
Colorectal MT-TT G15928A Ovary D-loop A16162G
Colorectal MT-CO1 C6371T Ovary MT-TW T5567C
Colorectal MT-ND5 T14138C Ovary D-loop C16527T
Colorectal MT-ND1 C3689G Pancreas MT-RNR2 G2946A
Colorectal MT-TR T10463C Pancreas MT-RNR2 G2145A
Colorectal MT-ND1 G3955A Pancreas D-loop G316A
Endometrial MT-CYB T15831C Prostate MT-ND3 A10398G
Endometrial MT-TW T5567C Prostate MT-TT G15928A
Endometrial MT-ND3 G10290A Prostate MT-TR T10463C
Endometrial MT-TR T10463C Prostate MT-CO3 G9820A
Endometrial MT-TP G15995A Prostate MT-ND5 A13966G
Glioblastoma D-loop C16134T Prostate D-loop G207A
Head & neck MT-ND2 T4823C Thyroid Intergenic A5581G
Leukemia MT-ND1 T3394C Thyroid D-loop G207A
Lung MT-ND2 T4823C Thyroid MT-TR T10463C
Lung MT-ND1 T3394C Thyroid MT-CO1 T7389C
Melanoma D-loop T310C
Melanoma D-loop A302CC
Melanoma D-loop T16519C
Nasopharyngeal D-loop C16261T
a

Shaded boxes indicate polymorphisms.

Table 2.

mtDNA haplotype predictions in cancer risk

Haplotype Ancestry/race Cancer tissue of origin Risk change
B2 American Cervix Increase
D Asian Endometrium Increase
D South American Cervix Increase
D Asian NSCLC Reduced survival
D Asian Lung Decrease
D Asian Esophagus Increase
D4 Asian Thyroid Increase
D5 Asian Thyroid Increase
D5 Asian Breast Increase
H European Breast Increase
I European Breast Increase
J Middle East Uveal Increase
K European Pancreas Decrease
K European Breast Increase
K European Thyroid Decrease
L0 African Prostate Increased severity
M Asian Oral Increase
M Asian Cervix Increase
M Asian Stomach Increase
M Asian Breast Decrease
M7 Asian Lung Increase
M7 Asian Liver Decrease
N European Prostate Decrease
N Asian Thyroid Increase
N Asian Stomach Longer survival
N Asian Breast Increase
N Asian Prostate Decrease
N Asian Breast Increase
T Multiple Colorectal Increase
U European Prostate Increase
U European Breast Decrease
U European Kidney Increase
X European Breast Increase

At the outset it is important to emphasize that findings regarding linkage of mtDNA SNPs and disparities are mostly anecdotal. Note, however, that we have shown that mtDNA-derived short RNA transcripts containing no SNPs are associated with ancestry or sex disparities in prostate [70], bladder [71], lung [71], kidney [71], and triple-negative breast cancer [72], and with survival in uveal melanoma patients [73]. A point worth emphasizing is that not all SNPs are within protein-coding mtDNA genes. Unfortunately, although examples exist, the literature exploring the other regions of the mitochondrial genome has not been as extensively studied so far.

Definitive cause–effect relationships are incomplete. Nonetheless, there are some intriguing observations and patterns that certainly warrant further investigation. An example includes SNPs found in the J haplotype which alter uveal cancer metastasis as well as mitochondrial biogenesis [13]. In addition, haplogroup N contributes to good survival in gastric carcinoma patients compared to haplogroup M, suggesting that SNPs in the displacement loop (D-loop) can be used as predictors of disease outcome [74]. Mutations in the D-loop are associated with age of onset, relapse rate, and metastasis for fibrous histiocytoma [75]. Polymorphisms in head and neck squamous cell carcinoma (HNSCC) are correlated with the mitochondrial unfolded protein response [76]. In a cohort of hepatocellular carcinomas, Li and colleagues [77] reported a large number of sequence variants in 140 cases, most of which were SNPs. The presence of these variants identified independent predictive factors in the D-loop for time to recurrence, tumor-free survival time, and overall survival. In another study, mitochondrial genomic variation in prostate cancers was greater at metastatic sites than in the primary tumor [78]. Relatedly, polymorphisms in the control region have been associated with melanoma etiology, pathogenesis, and progression [79].

Using cybrids, transfer of mtDNA SNPs found in breast cancer cells from African-American women slowed proliferation, increased complex I activity and ROS production, depolarized the mitochondria corresponding to apoptosis resistance, increased anchorage-independent growth, and increased metastases [80]. Likewise, mitochondria from an aggressive osteosarcoma cell line, a moderately aggressive breast carcinoma cell line, or an immortalized, but otherwise normal, breast epithelial cell demonstrated that mitochondria from the benign cells could reverse the cell proliferation, responses to hypoxia, apoptotic sensitivity, invasiveness, anchorage-independent growth, and tumorigenicity when injected into athymic mice [81]. In addition to the insights regarding tumor cell behavior, these findings suggest that mitochondrial gene therapy may be a viable option once deeper understanding is gained. Collectively, the results show that the mtDNA genome is not mutated randomly in metastases, suggesting that there were selective pressures for specific mutations associated with bone metastasis. Equally importantly, many of the SNPs are not in the protein-coding genes of the mtDNA. Both points will impinge upon future attempts to repair or restore normal phenotypes in cancer cells.

Because the presence of ROS is often associated with DNA damage, Blein et al. [82] and Ju et al. [25] measured mtDNA mutations in the presence or absence of DNA repair machinery. They reported that BRCA1/2 mutation carriers from the T haplogroup had an increased risk of developing breast cancer. By contrast, Ju et al. found no correlations between oncogenic drivers that impact on nDNA mutation rates and mtDNA mutation rates. Of 2453 colorectal cancers and 11 930 controls, the most significant mtDNA SNP was A4917G in the T haplogroup of American men and women of Asian, African, European, Latino, or Native Hawaiian ancestry [83]. This SNP was associated with an increased risk of developing colorectal cancer. Interestingly, European-Americans with G4655A were associated with higher colorectal cancer risk, but the effect was not observed throughout the population. Perhaps this SNP adds risk depending upon the nuclear factors with which it interacts.

Somatic mitochondrial mutations in cancer progression

Accumulating evidence suggests that mtDNA is correlated with cancer. Although many mitochondrial transcriptomic differences have been linked to disparities and survival, few driver mutations have been found. Some somatic mutations have been associated with specific cancer types (and subtypes) in both protein-coding and noncoding regions of the mitochondrial genome (reviewed in [8487]). Of note, some studies report a higher mutation frequency in the D-loop ([24,8890] and Table 1), but the underlying role(s) have not yet been determined. Another common deletion occurs in a stretch of cytosines (D310) that is variable. Similarly, correlations of mtDNA mutations or SNPs with metastasis have been observed, but ascribing a mechanism to those changes has not been possible to date. This situation is also true in the context of metastasis, but some mtDNA changes associated with oncogenesis will also likely provide insights into the ontogeny and/or the severity of metastasis.

In clear cell renal carcinoma, Zhang et al. observed a generalized decrease of complex I despite relatively few mutations in complex I proteins [58]. The expression patterns appeared to correlate with development of metastasis and immune response. Zhan and colleagues observed a novel tRNA fragment in the serum of hepatocellular carcinoma patients that had predictive value [91]. In animal models of liver cancer, c-Myc-driven mtDNA fragmentation was measured as a possible driver of tumor formation. Knockout of the mitochondrial fusion molecule MFN1 or overexpression of mitochondrial fission molecule DRP1 promoted liver cancer development [68]. The opposite effect was observed when MFN1 was overexpressed.

Mitochondria are closely associated with the actin cytoskeleton and are highly mobile within cells. Recent elegant studies by Caino and colleagues demonstrate that mitochondria mobilization in cancer cells co-opts a neuronal mitochondrial traffic machinery involving syntaphilin (SNPH), kinesin KIF5B, and GTPase Miro1/2 to localize mitochondria to the cortical cytoskeleton [92,93]. Overexpression of SNPH suppressed the kinetics and distance traveled by mitochondria, with a corresponding reduction in chemotaxis and metastasis in vivo. More extensive analyses reveal that mitochondrial subcellular localization is dysregulated in multiple cancer cells [94,95]. However, another question arises: are there differences in mtDNA subcellular distribution in different primary and metastatic lesions of different cancers, or between different sites of metastasis?

Mitochondrial genetics and metastasis

Metastasis is a highly inefficient process that is responsible for the majority of cancer-related morbidity and mortality, and is a culmination of cellular evolution from transformation to the neoplastic state [96,97]. Both pro- and antimetastatic genes have been discovered, but there are many pathways to achieve metastatic colonization [97]. Reviewing the literature reveals several suggestive links between mitochondria and metastasis, but definitive roles still require experimental testing. A common conclusion is that mitochondrial genetic changes are phenotypic ‘modifiers’ rather than drivers per se.

Mitochondrial haplogroups have been implicated in multiple hallmarks of metastasis [44]. For example, native American haplogroups in Mexican women with breast cancer showed no statistical difference between haplogroups A–D, and L according to tumor subtype [42] whereas in-depth analysis uncovered many important results. Single-nucleotide variants comprised 98% of somatic mutations; 59% of total mutations were in coding regions and 41% were in noncoding regions. The D-loop had the most mutations, followed by CO1, 16S, 12S, and ND5. Somatic mutations were homoplasmic (21.4%), but most were heteroplasmic.

Invasive breast cancers in African-American women participating in the University of North Carolina Breast Cancer Study had increased risk with G10398A, which had previously been linked to neurodegenerative diseases [98]. Curiously, Caucasian women harboring the same allele showed no difference in relative risk. These findings affirm the cooperativity between genomes as QTLs. It is interesting to speculate, based upon the higher frequency of developing triple-negative breast cancer in African-American women, that part of these differences could be due to disruption of estrogen-based regulation of mitochondrial function [99,100]. However, there are no direct data that establish such a link.

Most data aforementioned describe mutations in mtDNA coding regions. Webb et al. reported no correlation between overall colorectal cancer risk and mitochondrial haplogroup [101]. They did, however, find that A5657G in the non-coding region between mt-tRNAAla and mt-tRNAAsp correlated with colorectal (compared to rectal) cancers. A synonymous change in MT-ND2 (T4562C) was highly associated with microsatellite instability of colorectal cancers [101].

Hunter and colleagues studied genetic background as a metastasis efficiency-modifier loci by crossing FVB/NJ-TgN (MMTVPyMT)634nul mice (‘PyMT’) to dams from different mouse strains. F1 progeny had varied metastatic efficacy [102105]. They have since mapped several metastasis efficiency-modifiers in breast and prostate cancers [102,104,106,107]. However, their experimental design allowed us to posit an alternative interpretation based upon maternal inheritance [108]. We developed a genetically engineered mouse model by transferring pronuclei from one mouse strain into the enucleated embryonic cytoplast of another strain [109,110]. Following transfer into pseudopregnant nuclear-matched mice, pups were designated mitochondrial-nuclear exchange (MNX) mice. We replicated a scaled-down version of the Hunter experiment. Changes in tumor incidence, latency, and metastasis were nearly indistinguishable from those using wild-type mice [108,111,112]. Comparable changes were observed using HER2/neu as an oncogenic driver [111]. MNX mice also exhibited differences in spontaneous tumor formation even in the absence of an oncogenic driver [113] in addition to differences in atherosclerosis, adiposity, and diabetes [110,114117].

Recognizing that genetic crosses also change stromal mtDNA, we asked whether syngeneic (i.e., histocompatible) mammary and melanoma tumor cells behaved differently when injected into MNX mice. Whenever host stroma contained C57BL/6J mtDNA, metastasis was inhibited; whenever stroma had C3H/HeN mtDNA, metastasis was promoted [118]. In efforts to understand the underlying mechanisms, we determined that MNX mice selectively alter nuclear DNA methylation and histone marks ([112]; X. McGuire and D.R.W., unpublished). C3H/HeN mtDNA consistently produces higher levels of ROS than C57BL/6J mtDNA. Metastasis increased with higher ROS whereas scavenging ROS decreased metastases [111]. Baseline immune profiles differed in MNX mice compared to nDNA-matched counterparts, and the polarization states of myeloid populations infiltrating lung metastases were also different [119]. Principal component analyses of >5000 metabolites in wild-type and MNX mice clustered slightly differently but were largely similar (D.R.W., unpublished). Specific metabolites corresponding to changing metastatic potential have not yet been identified, consistent with previous reports [110,111,120123]. Because mitochondria evolved from ancient bacteria [124], and because bacterial ecosystems form when microbes communicate with each other [125127], we reasoned that mitochondria could communicate with some bacteria, thereby selectively promoting/inhibiting bacterial growth and fecal microbiome change. Deep sequencing of 16S RNA identified selective changes (only 8–17 bacterial species) (D.R.W. et al. unpublished).

MNX mouse findings mice support pioneering work from Ishikawa, who utilized cybrids to demonstrate that mitochondrial transfer could change metastasis [61,128,129]. Mutations disrupting complex I (e.g., an insertion into the MT-ND6 gene, 13885insC) elevated ROS and metastatic propensity [61,62]. mtDNA mutation also increased the transcription of glycolysis- and metastasis-related genes [62]. The functional role of complex I mutations in MT-ND6 (C12084T) and MT-ND5 (A13966G) was later linked to metastasis [129]. Mutant MT-ND6 increased invasion in A549 lung cancer cells [130]. Multiple mutations in NADH dehydrogenase genes (T3398C, T12338C, C3689G, G3709A, G3955A, T10363C, C11409T, G13103A, and T14138CC in MT-ND1; G12813A, G13366A, and 14504delA premature truncations in MT-ND5 or MT-ND6) were associated with distant metastasis [62,131,132]. Two SNPs were found in MT-ND1 (C3497T and T3394C), consistent with the notion of ancestral disparities in cancer aggressiveness. Paralleling our observations that mtDNA SNPs in stroma affect metastasis in MNX mice, T3394C mutations in adjacent mucosa in NSCLC and colon tumors [62] support the concept that there are inherited susceptibilities to metastasis. Comparison of non-invasive versus invasive breast cancer cell lines observed shifts from OXPHOS to glycolysis together with increased heteroplasmy [133]. However, such patterns are not universally reported in all cancer types [46,134]. Other reports identified antioxidants (e.g., mitochondrial catalase, mtCAT) that affect metastasis [135,136]. mtCAT reduced macrophage infiltration and CD34+ endothelial cells, the latter suggesting reduced angiogenesis.

The strongest evidence for germline metastasis efficiency loci in mtDNA comes from experimental models, most commonly mice. Although powerful, mouse models have weaknesses (reviewed by Bussard and Siracusa [137]). Indeed, extrapolating murine data to humans is limited by incomplete tools and the limited number of haplogroups, making it difficult to separate nuclear and mitochondrial contributions. However, recent publications implicate mitochondrial haplotypes as modifiers of hepatocellular carcinoma [138] and neuroblastoma [139].

Most examples focus on mtDNA changes in the cancer cells themselves. Understanding the involvement of mtDNA in the stroma is complicated by variances in tissue, gender, age, and stress. The experiments cited below are individually well controlled, but current technology and experimental costs make it impractical to address all these variables.

Immune activation and mitochondria functions are inextricably linked. Mitochondria regulate innate immune system activation through recognition of cellular damage and activation of the NLRP3 inflammasome [140,141], and formylated peptides in bacteria can also activate specific receptors [142]. In addition, ROS production can indirectly or directly regulate the immune system [143,144]. Cancer-associated fibroblasts (CAFs) can produce pyruvate, lactate, ketone bodies, and fatty acids using glycolysis that can be utilized by cancer cells [145151]. These same metabolites also alter immune cell functions [152] and polarization [153]. As the primary consumers of oxygen in most tissues, mitochondrial differences in oxygenation regulate metabolism [154]. Mitochondria-derived ROS can induce hypoxia-inducible factor α (HIF-1α) which in turn alters cellular responses [155]. Thus, hypoxia will have overarching effects on many cells within the tumor microenvironment, which may be further impacted by differences in mtDNA. Because hypoxia increases metastasis [156160], the capacity of mitochondria to respond may be intrinsic to connecting the two phenotypes.

Concluding remarks and future perspectives

Despite limited models that isolate mitochondrial SNPs and mitochondrial genetics as experimental variables, abundant data support the notion that mtDNA QTLs exist for multiple diseases including cancer. It is unfortunate that most investigators default to ignoring the mitochondrial genome because of its relatively small size, and others focus solely on the protein-coding regions of the mitochondrial genome. We remind readers that small size does not always mean little impact and that ‘junk’ DNA is now well recognized as a crucial regulator of cellular function. The influence of mitochondria on cancer development, progression, and disease severity occurs because of both intrinsic and extrinsic effects. As conveyers of signals from the microenvironment to the nucleus and vice versa, mitochondrial SNPs can alter signaling in context-dependent ways. What are the signals? Are the same molecules used to signal multiple locations? With what are those molecules interacting? How do nuclear genes and the proteins (and non-coding RNA) derived from them modify the mtDNA?

Where should we go from here? Hopefully, this review overcomes the misconception that the mitochondrial genome is (relatively) unimportant. Throughout this review we have attempted to introduce questions and topics for future research. Some of the bigger issues are summarized in the Outstanding questions. The most important recommendation is to increase the incorporation of mtDNA analyses in cancer genomic analyses. When doing so, pairing normal, primary tumor, and metastatic tissues will be crucial, especially from different metastatic sites.

Because germline differences in mtDNA perhaps predict tumor behavior, and because ostensibly normal tissues adjacent to diseased tissues are not truly normal [161], there is immense potential that mitochondrial information within serum, plasma, or non-tumor tissues will be predictive or prognostic, even before oncogenic mutations appear or before tumors arise [73]. Similar findings have been described for other human conditions [162]. The relatively small size of the mitochondrial genome could, in fact, be a benefit because deep sequencing would be comparatively fast and inexpensive compared to whole-genome or even whole-exome sequencing.

Another crucial step will be to develop databases that capture mitochondrial genomic properties fully and effectively allow comparisons of non-human with human mtDNA. Furthermore, the development of technologies that allow systematic manipulation of the mitochondrial genome without introducing heteroplasmy [163] will accelerate the study of mitochondrial genomics. The field of mitochondrial cancer genetics is still in its relative infancy. Nonetheless, the foundations have been laid. New and emerging models hold great promise for addressing these questions and initiating the pathway to translating the foundational questions into clinical practice.

Table I.

Organization of human and murine mitochondrial genomes

Human Murine
Genea ETCb Start End Length (bp) Strand (H or L) Start End Length (bp) Strand (H or L)
MT-TF 577 647 70 H 1 68 67 H
MT-RNR1 648 1601 953 H 70 1024 954 H
MT-TV 1602 1670 68 H 1025 1093 68 H
MT-RNR2 1671 3229 1558 H 1094 2675 1581 H
MT-TL1 3230 3304 74 H 2676 2750 74 H
MT-ND1 I 3307 4262 955 H 2760 3704 944 H
MT-TI 4263 4331 68 H 3706 3774 68 H
MT-TQ 4329 4400 71 L 3772 3842 70 L
MT-TM 4402 4469 67 H 3845 3913 68 H
MT-ND2 I 4470 5511 1041 H 3914 4948 1034 H
MT-TW 5512 5579 67 H 4950 5016 66 H
MT-TA 5587 5655 68 L 5018 5086 68 L
MT-TN 5657 5729 72 L 5089 5159 70 L
MT-TC 5761 5826 65 L 5193 5257 64 L
MT-TY 5826 5891 65 L 5260 5326 66 L
MT-CO1 III/IV 5904 7445 1541 H 5328 6872 1544 H
MT-TS1 7446 7514 68 L 6869 6939 70 L
MT-TD 7518 7585 67 H 6942 7011 69 H
MT-CO2 IV 7586 8269 683 H 7013 7696 683 H
MT-TK - 8295 8364 69 H 7700 7764 64 H
MT-ATP8 V 8366 8572 206 H 7766 7969 203 H
MT-ATP6 V 8527 9207 680 H 7927 8607 680 H
MT-CO3 IV 9207 9990 783 H 8606 9389 783 H
MT-TG - 9991 10058 67 H 9391 9458 67 H
MT-ND3 I 10059 10404 345 H 9457 9803 346 H
MT-TR 10405 10469 64 H 9805 9872 67 H
MT-ND4L I 10470 10766 296 H 9874 10167 293 H
MT-ND4 I 10760 12137 1377 H 10161 11537 1376 H
MT-TH 12138 12206 68 H 11539 11606 67 H
MT-TS2 12207 12265 58 H 11607 11665 58 H
MT-TL2 12266 12336 70 H 11665 11735 70 H
MT-ND5 I 12337 14148 1811 H 11736 13559 1823 H
MT-ND6 I 14149 14673 524 L 13546 14064 518 L
MT-TE 14674 14742 68 L 14065 14133 68 L
MT-CYB III 14747 15887 1140 H 14139 15281 1142 H
MT-TT 15888 15953 65 H 15283 15449 166 H
MT-TP 15956 16023 67 L 15350 15416 66 L
D-loop 16024 576 1121 15417 16295 878
a

Gene names are preceded by ‘MT’ in accordance with Human Genome Organization (HUGO) Gene Nomenclature Committee recommendations.

b

Indicates whether the designated gene encodes a component of electron transport chain (ETC) complexes I–IV; –, does not encode an ETC component.

Highlights.

Mitochondrial DNA (mtDNA) harbors cancer/metastasis quantitative trait loci.

Both somatic mutations and germline mtDNA polymorphisms are associated with cancer development and metastasis in tissue-specific and cancer subtype-specific manners.

Mitochondria alter nuclear epigenomes.

Not all mitochondrial effects on cancer/metastasis are metabolism (i.e., electron transport)-related.

Changes in stromal mitochondria influence neoplastic behavior.

Acknowledgments

Work in the laboratory of D.R.W. has been generously funded by Susan G. Komen for the Cure (SAC110037), the National Foundation for Cancer Research, METAvivor Research and Services Inc., and Theresa’s Research Foundation. Additional funding support was provided by US Army Medical Research Defense Command (W81XWH-18-1-0450) and National Cancer Institute (P30-CA168524). We apologize to any authors whose work was omitted due to article guidelines.

Glossary

Cybrid

a fused cell containing a parental nuclear genome and the mitochondrial genome of another parent (or a combination of mitochondrial genomes)

Haplotype

alleles inherited from a single parent

Haplogroup

a clade in which a unique polymorphism is represented

Heteroplasmy/heteroplasmic

a eukaryotic cell in which mitochondrial DNA (mtDNA) is non-identical

Homoplasmy/homoplasmic

a eukaryotic cell in which all copies of mtDNA are identical

Quantitative trait loci (QTLs)

genes with alleles that influence the expression of a phenotypic trait

Metastasis

the spread of neoplastic cells to discontiguous sites where neoplastic cells proliferate to become macroscopic lesions

Metastatic cascade

the series of steps used by cancer cells undergoing the process of metastasis

Tunneling nanotubes (TNTs)

cellular protrusions that enable cells to touch over long distances

Footnotes

Declaration of interests

The authors declare no conflicts of interest.

References

  • 1.Clyde D (2022) Mitochondrial DNA copy number and disease. Nat. Rev. Genet 23, 136. [DOI] [PubMed] [Google Scholar]
  • 2.Chong M et al. (2022) GWAS and ExWAS of blood mitochondrial DNA copy number identifies 71 loci and highlights a potential causal role in dementia. eLife 11, e70382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Nguyen NNY et al. (2020) Deregulated mitochondrial DNA in diseases. DNA Cell Biol. 39, 1385–1400 [DOI] [PubMed] [Google Scholar]
  • 4.Linnane AW et al. (1989) Mitochondrial DNA mutations as an important contributor to ageing and degenerative diseases. Lancet 1, 642–645 [DOI] [PubMed] [Google Scholar]
  • 5.Szczepanowska K and Trifunovic A (2017) Origins of mtDNA mutations in ageing. Essays Biochem. 61, 325–337 [DOI] [PubMed] [Google Scholar]
  • 6.Sharpley MS et al. (2012) Heteroplasmy of mouse mtDNA is genetically unstable and results in altered behavior and cognition. Cell 151, 333–343 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Yapa NMB et al. (2021) Mitochondrial dynamics in health and disease. FEBS Lett. 595, 1184–1204 [DOI] [PubMed] [Google Scholar]
  • 8.Brown JA et al. (2020) An evolutionary, or ‘mitocentric’ perspective on cellular function and disease. Redox Biol. 36, 101568. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Schon KR et al. (2020) Mitochondrial diseases: a diagnostic revolution. Trends Genet. 36, 702–717 [DOI] [PubMed] [Google Scholar]
  • 10.Taylor RW and Turnbull DM (2005) Mitochondrial DNA mutations in human disease. Nat. Rev. Genet 6, 389–402 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Silva-Pinheiro P and Minczuk M (2022) The potential of mitochondrial genome engineering. Nat. Rev. Genet 23, 199–214 [DOI] [PubMed] [Google Scholar]
  • 12.Russell OM et al. (2020) Mitochondrial diseases: hope for the future. Cell 181, 168–188 [DOI] [PubMed] [Google Scholar]
  • 13.Singh L et al. (2021) Mitochondrial DNA polymorphisms and biogenesis genes in primary and metastatic uveal melanoma cell lines. Cancer Genet. 256, 91–99 [DOI] [PubMed] [Google Scholar]
  • 14.Perez-Amado CJ et al. (2021) Mitochondrial heteroplasmy shifting as a potential biomarker of cancer progression. Int. J. Mol. Sci 22, 25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Reznik E et al. (2017) Mitochondrial respiratory gene expression is suppressed in many cancers. eLife 6, e21592. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Kalsbeek AMF et al. (2017) Mitochondrial genome variation and prostate cancer: a review of the mutational landscape and application to clinical management. Oncotarget 8, 71342–71357 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Reznik E et al. (2016) Mitochondrial DNA copy number variation across human cancers. eLife 5, e10769. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Hsu CC et al. (2016) Role of mitochondrial dysfunction in cancer progression. Exp. Biol. Med. (Maywood) 241, 1281–1295 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.van Gisbergen MW et al. (2015) How do changes in the mtDNA and mitochondrial dysfunction influence cancer and cancer therapy? Challenges, opportunities and models. Mutat. Res. Rev. Mutat. Res 764, 16–30 [DOI] [PubMed] [Google Scholar]
  • 20.Chan DC (2020) Mitochondrial dynamics and its involvement in disease. Annu. Rev. Pathol 15, 235–259 [DOI] [PubMed] [Google Scholar]
  • 21.Zhu D et al. (2022) Mitochondrial-to-nuclear communication in aging: an epigenetic perspective. Trends Biochem Sci. Published online April 6, 2022. 10.1016/j.tibs.2022.03.008 [DOI] [PubMed] [Google Scholar]
  • 22.Matilainen O et al. (2017) Mitochondria and epigenetics – crosstalk in homeostasis and stress. Trends Cell Biol. 27, 453–463 [DOI] [PubMed] [Google Scholar]
  • 23.Picard M et al. (2014) Progressive increase in mtDNA 3243A>G heteroplasmy causes abrupt transcriptional reprogramming. Proc. Natl. Acad. Sci. U. S. A 111, E4033–E4042 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Stewart JB et al. (2015) Simultaneous DNA and RNA mapping of somatic mitochondrial mutations across diverse human cancers. PLoS Genet. 11, e1005333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Ju YS et al. (2014) Origins and functional consequences of somatic mitochondrial DNA mutations in human cancer. eLife 3, e02935. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Sercel AJ et al. (2021) Mitochondrial DNA dynamics in reprogramming to pluripotency. Trends Cell Biol. 31, 311–323 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Warburg O et al. (1927) The metabolism of tumors in the body. J. Gen. Physiol 8, 519–530 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Warburg O (1956) On respiratory impairment in cancer cells. Science 124, 269–270 [PubMed] [Google Scholar]
  • 29.Warburg O (1956) On the origin of cancer cells. Science 123, 309–314 [DOI] [PubMed] [Google Scholar]
  • 30.Crabtree HG (1929) Observations on the carbohydrate metabolism of tumours. Biochem. J 23, 536–545 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Lunt SY and Vander Heiden MG (2011) Aerobic glycolysis: meeting the metabolic requirements of cell proliferation. Annu. Rev. Cell Dev. Biol 27, 441–464 [DOI] [PubMed] [Google Scholar]
  • 32.Yang H et al. (2012) IDH1 and IDH2 mutations in tumorigenesis: mechanistic insights and clinical perspectives. Clin. Cancer Res 18, 5562–5571 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Loureiro R et al. (2017) Mitochondrial biology in cancer stem cells. Semin. Biol 47, 18–28 [DOI] [PubMed] [Google Scholar]
  • 34.Frezza C and Gottlieb E (2009) Mitochondria in cancer: not just innocent bystanders. Semin. Cancer Biol 19, 4–11 [DOI] [PubMed] [Google Scholar]
  • 35.Choudhury AR and Singh KK (2017) Mitochondrial determinants of cancer health disparities. Semin. Cancer Biol 47, 125–146 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Chakrabarty S et al. (2018) Mitochondria in health and disease. Mitochondrion 43, 25–29 [DOI] [PubMed] [Google Scholar]
  • 37.Yuan Y et al. (2020) Comprehensive molecular characterization of mitochondrial genomes in human cancers. Nat. Genet 52, 342–352 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Correia M et al. (2017) Etiopathogenesis of oncocytomas. Semin. Cancer Biol 47, 82–94 [DOI] [PubMed] [Google Scholar]
  • 39.Zhang W et al. (2018) A global transcriptional network connecting noncoding mutations to changes in tumor gene expression. Nat. Genet 50, 613–620 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Cookson W et al. (2009) Mapping complex disease traits with global gene expression. Nat. Rev. Genet 10, 184–194 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Hunter KW et al. (2018) Genetic insights into the morass of metastatic heterogeneity. Nat. Rev. Cancer 18, 211–223 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Perez-Amado CJ et al. (2020) Mitochondrial DNA mutation analysis in breast cancer: shifting from germline heteroplasmy toward homoplasmy in tumors. Front. Oncol 10, 572954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Scheid AD et al. (2019) The second genome: effects of the mitochondrial genome on cancer progression. Adv. Cancer Res 142, 63–105 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Scheid AD et al. (2021) Roles of mitochondria in the hallmarks of metastasis. Br. J. Cancer 124, 124–135 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Wang P et al. (2021) Epigenome-wide association study of mitochondrial genome copy number. Hum. Mol. Genet 31, 309–319 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Liu W et al. (2014) Metastasis suppressor KISS1 seems to reverse the Warburg effect by enhancing mitochondrial biogenesis. Cancer Res. 74, 954–963 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Sahu P et al. (2018) Tunneling nanotubes: a versatile target for cancer therapy. Curr. Cancer Drug Targets 18, 514–521 [DOI] [PubMed] [Google Scholar]
  • 48.Allegra A et al. (2022) Specialized intercellular communications via tunnelling nanotubes in acute and chronic leukemia. Cancers (Basel) 14, 21. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Takenaga K et al. (2021) Intercellular transfer of mitochondrial DNA carrying metastasis-enhancing pathogenic mutations from high- to low-metastatic tumor cells and stromal cells via extracellular vesicles. BMC Mol. Cell Biol 22, 52. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Zampieri LX et al. (2021) Mitochondrial transfer in cancer: a comprehensive review. Int. J. Mol. Sci 22, 3245. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Hekmatshoar Y et al. (2018) The role of metabolism and tunneling nanotube-mediated intercellular mitochondria exchange in cancer drug resistance. Biochem. J 475, 2305–2328 [DOI] [PubMed] [Google Scholar]
  • 52.Lu J et al. (2017) Tunneling nanotubes promote intercellular mitochondria transfer followed by increased invasiveness in bladder cancer cells. Oncotarget 8, 15539–15552 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Marlein CR et al. (2019) CD38-driven mitochondrial trafficking promotes bioenergetic plasticity in multiple myeloma. Cancer Res. 79, 2285–2297 [DOI] [PubMed] [Google Scholar]
  • 54.Wang X and Gerdes HH (2015) Transfer of mitochondria via tunneling nanotubes rescues apoptotic PC12 cells. Cell Death Differ. 22, 1181–1191 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Pasquier J et al. (2013) Preferential transfer of mitochondria from endothelial to cancer cells through tunneling nanotubes modulates chemoresistance. J. Transl. Med 11, 94. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Kheirandish-Rostami M et al. (2020) Mitochondrial characteristics contribute to proliferation and migration potency of MDA-MB-231 cancer cells and their response to cisplatin treatment. Life Sci. 244, 117339. [DOI] [PubMed] [Google Scholar]
  • 57.Derdak Z et al. (2008) The mitochondrial uncoupling protein-2 promotes chemoresistance in cancer cells. Cancer Res. 68, 2813–2819 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Zhang FT et al. (2022) Comprehensive analysis of lower mitochondrial complex I expression is associated with cell metastasis of clear cell renal cell carcinoma. Transl. Cancer Res 16, 1488–1502 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Urra FA et al. (2017) The mitochondrial complex(I)ty of cancer. Front. Oncol 7, 118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Santidrian AF et al. (2013) Mitochondrial complex I activity and NAD+/NADH balance regulate breast cancer progression. J. Clin. Invest 123, 1068–1081 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Ishikawa K et al. (2008) ROS-generating mitochondrial DNA mutations can regulate tumor cell metastasis. Science 320, 661–664 [DOI] [PubMed] [Google Scholar]
  • 62.Koshikawa N et al. (2017) Association of predicted pathogenic mutations in mitochondrial ND genes with distant metastasis in NSCLC and colon cancer. Sci. Rep 7, 15535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Yokota M et al. (2010) Generation of trans-mitochondrial mitomice by the introduction of a pathogenic G13997A mtDNA from highly metastatic lung carcinoma cells. FEBS Lett. 584, 3943–3948 [DOI] [PubMed] [Google Scholar]
  • 64.Boroughs LK and DeBerardinis RJ (2015) Metabolic pathways promoting cancer cell survival and growth. Nat. Cell Biol 17, 351–359 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Dang CV (2010) Rethinking the Warburg effect with Myc micromanaging glutamine metabolism. Cancer Res. 70, 859–862 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Cantley LC et al. (1991) Oncogenes and signal transduction. Cell 64, 281–302 [DOI] [PubMed] [Google Scholar]
  • 67.Basu S et al. (2018) Mutant p53 controls tumor metabolism and metastasis by regulating PGC-1alpha. Genes Dev. 32, 230–243 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Wang D et al. (2022) Mitochondrial fragmentation is crucial for c-Myc-driven hepatoblastoma-like liver tumors. Mol. Ther 30, 1645–1660 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Beadnell TC et al. (2018) Roles of the mitochondrial genetics in cancer metastasis: not to be ignored any longer. Cancer Metastasis Rev. 37, 615–632 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Magee RG et al. (2018) Profiles of miRNA isoforms and tRNA fragments in prostate cancer. Sci. Rep 8, 5314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Telonis AG et al. (2019) tRNA fragments show intertwining with mRNAs of Specific repeat content and have links to disparities. Cancer Res. 79, 3034–3049 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Telonis AG and Rigoutsos I (2018) Race disparities in the contribution of miRNA isoforms and tRNA-derived fragments to triple-negative breast cancer. Cancer Res. 78, 1140–1154 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Londin E et al. (2020) IsomiRs and tRNA-derived fragments are associated with metastasis and patient survival in uveal melanoma. Pigment Cell Melanoma Res. 33, 52–62 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Wang C et al. (2014) Mitochondrial DNA haplogroup N is associated good outcome of gastric cancer. Tumour Biol. 35, 12555–12559 [DOI] [PubMed] [Google Scholar]
  • 75.Luo ZC et al. (2019) Single nucleotide polymorphisms in the D-loop region predicts earlyage-at-onset of malignant fibrous histiocytoma. Mitochondrial. DNA B Resour 4, 2078–2083 [Google Scholar]
  • 76.Ahmed MW et al. (2019) Relationship of single nucleotide polymorphisms and haplotype interaction of mitochondrial unfolded protein response pathway genes with head and neck cancer. Future Oncol. 15, 3819–3829 [DOI] [PubMed] [Google Scholar]
  • 77.Li S et al. (2016) Associations between sequence variations in the mitochondrial DNA D-loop region and outcome of hepatocellular carcinoma. Oncol. Lett 11, 3723–3728 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Arnold RS et al. (2015) Bone metastasis in prostate cancer: recurring mitochondrial DNA mutation reveals selective pressure exerted by the bone microenvironment. Bone 78, 81–86 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Ebner S et al. (2011) Mitochondrial haplogroups, control region polymorphisms and malignant melanoma: a study in middle European Caucasians. PLoS One 6, e27192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Kulawiec M et al. (2009) mtDNA G10398A variant in African-American women with breast cancer provides resistance to apoptosis and promotes metastasis in mice. J. Hum. Genet 54, 647–654 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Kaipparettu BA et al. (2013) Crosstalk from non-cancerous mitochondria can inhibit tumor properties of metastatic cells by suppressing oncogenic pathways. PLoS One 8, e61747. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Blein S et al. (2015) Targeted sequencing of the mitochondrial genome of women at high risk of breast cancer without detectable mutations in BRCA1/2. PLoS One 10, e0136192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Li Y et al. (2015) Association of genes, pathways, and haplogroups of the mitochondrial genome with the risk of colorectal cancer: the Multiethnic Cohort. PLoS One 10, e0136796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Chatterjee A et al. (2006) Mitochondrial DNA mutations in human cancer. Oncogene 25, 4663–4674 [DOI] [PubMed] [Google Scholar]
  • 85.Lu J et al. (2009) Implications of mitochondrial DNA mutations and mitochondrial dysfunction in tumorigenesis. Cell Res. 19, 802–815 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Hertweck KL and Dasgupta S (2017) The landscape of mtDNA modifications in cancer: a tale of two cities. Front. Oncol 7, 262. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Brandon MC et al. (2005) MITOMAP: a human mitochondrial genome database–2004 update. Nucleic Acids Res. 33, D611–D613 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Akouchekian M et al. (2009) High rate of mutation in mitochondrial DNA displacement loop region in human colorectal cancer. Dis. Colon Rectum 52, 526–530 [DOI] [PubMed] [Google Scholar]
  • 89.Bragoszewski P et al. (2008) Limited clinical relevance of mitochondrial DNA mutation and gene expression analyses in ovarian cancer. BMC Cancer 8, 292. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Parrella P et al. (2003) Mutations of the D310 mitochondrial mononucleotide repeat in primary tumors and cytological specimens. Cancer Lett. 190, 73–77 [DOI] [PubMed] [Google Scholar]
  • 91.Zhan S et al. (2022) Serum mitochondrial tsRNA serves as a novel biomarker for hepatocarcinoma diagnosis. Front. Med 16, 216–226 [DOI] [PubMed] [Google Scholar]
  • 92.Caino MC et al. (2016) A neuronal network of mitochondrial dynamics regulates metastasis. Nat. Commun 7, 13730. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Caino MC et al. (2017) Syntaphilin controls a mitochondrial rheostat for proliferation-motility decisions in cancer. J. Clin. Invest 127, 3755–3769 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Furnish M and Caino MC (2020) Altered mitochondrial trafficking as a novel mechanism of cancer metastasis. Cancer Rep. (Hoboken) 3, e1157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Yadav T et al. (2022) Mitochondria–actin cytoskeleton crosstalk in cell migration. J. Cell. Physiol 237, 2387–2403 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Eccles SA and Welch DR (2007) Metastasis: recent discoveries and novel treatment strategies. Lancet 369, 1742–1757 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Welch DR and Hurst DR (2019) Defining the hallmarks of metastasis. Cancer Res. 79, 3011–3027 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Canter JA et al. (2005) Mitochondrial DNA G10398A polymorphism and invasive breast cancer in African-American women. Cancer Res. 65, 8028–8033 [DOI] [PubMed] [Google Scholar]
  • 99.Klinge CM (2020) Estrogenic control of mitochondrial function. Redox Biol. 31, 101435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Klinge CM (2008) Estrogenic control of mitochondrial function and biogenesis. J. Cell. Biochem 105, 1342–1351 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Webb E et al. (2008) Comprehensive analysis of common mitochondrial DNA variants and colorectal cancer risk. Br. J. Cancer 99, 2088–2093 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Hunter KW (2012) Mouse models of cancer: does the strain matter? Nat. Rev. Cancer 12, 144–149 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Le Voyer T et al. (2000) An epistatic interaction controls the latency of a transgene-induced mammary tumor. Mamm. Genome 11, 883–889 [DOI] [PubMed] [Google Scholar]
  • 104.Le Voyer T et al. (2001) Three loci modify growth of a transgene-induced mammary tumor: suppression of proliferation associated with decreased microvessel density. Genomics 74, 253–261 [DOI] [PubMed] [Google Scholar]
  • 105.Lifsted T et al. (1998) Identification of inbred mouse strains harboring genetic modifiers of mammary tumor age of onset and metastatic progression. Int. J. Cancer 77, 640–644 [DOI] [PubMed] [Google Scholar]
  • 106.Winter JM et al. (2017) Mapping complex traits in a diversity outbred F1 mouse population identifies germline modifiers of metastasis in human prostate cancer. Cell Syst. 4, 31–45 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Ross C et al. (2020) Metastasis-specific gene expression in autochthonous and allograft mouse mammary tumor models: stratification and identification of targetable signatures. Mol. Cancer Res 18, 1278–1289 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Feeley KP et al. (2015) Mitochondrial genetics regulate breast cancer tumorigenicity and metastatic potential. Cancer Res. 75, 4429–4436 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Kesterson RA et al. (2016) Generation of mitochondrial–nuclear exchange mice via pronuclear transfer. Bio Protoc. 6, e1976. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Fetterman JL et al. (2013) Mitochondrial genetic background modulates bioenergetics and susceptibility to acute cardiac volume overload. Biochem. J 455, 157–167 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Brinker AE et al. (2017) Mitochondrial haplotype alters mammary cancer tumorigenicity and metastasis in an oncogenic driver-dependent manner. Cancer Res. 77, 6941–6949 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Vivian CJ et al. (2017) Mitochondrial genomic backgrounds affect nuclear DNA methylation and gene expression. Cancer Res. 77, 6202–6214 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Vivian CJ et al. (2018) Mitochondrial polymorphisms contribute to aging phenotypes in MNX mouse models. Cancer Metastasis Rev. 37, 633–642 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Kandasamy J et al. (2019) Mitochondrial DNA variation modulates alveolar development in newborn mice exposed to hyperoxia. Am. J. Physiol. Lung Cell Mol. Physiol 317, L740–L747 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Betancourt AM et al. (2014) Mitochondrial–nuclear genome interactions in non-alcoholic fatty liver disease in mice. Biochem. J 461, 223–232 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Dunham-Snary KJ et al. (2018) Mitochondrial–nuclear genetic interaction modulates whole body metabolism, adiposity and gene expression in vivo. EBioMedicine 36, 316–328 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Bray AW and Ballinger SW (2017) Mitochondrial DNA mutations and cardiovascular disease. Curr. Opin. Cardiol 32, 267–274 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Brinker AE et al. (2020) Mitochondrial haplotype of the host stromal microenvironment alters metastasis in a non-cell autonomous manner. Cancer Res. 80, 1118–1129 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Beadnell TC et al. (2020) Mitochondrial genetics cooperate with nuclear genetics to selectively alter immune cell development/trafficking. Biochim. Biophys. Acta Mol. basis Dis 1866, 165648. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Yu D et al. (2020) Mitochondrial metabolism and cancer metastasis. Ann. Transl. Med 8, 904. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Tiedemann K et al. (2020) Role of altered metabolic microenvironment in osteolytic metastasis. Front. Cell Dev. Biol 8, 435. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Ly T et al. (2020) KISS1 in metastatic cancer research and treatment: potential and paradoxes. Cancer Metastasis Rev. 39, 739–754 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Tasdogan A et al. (2020) Metabolic heterogeneity confers differences in melanoma metastatic potential. Nature 577, 115–120 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Wallace DC (2016) Mitochondrial DNA in evolution and disease. Nature 535, 498–500 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Bassler BL (1999) How bacteria talk to each other: regulation of gene expression by quorum sensing. Curr. Opin. Microbiol 2, 582–587 [DOI] [PubMed] [Google Scholar]
  • 126.Eickhoff MJ and Bassler BL (2018) SnapShot: bacterial quorum sensing. Cell 174, 1328. [DOI] [PubMed] [Google Scholar]
  • 127.Irie Y and Parsek MR (2008) Quorum sensing and microbial biofilms. Curr. Top. Microbiol. Immunol 322, 67–84 [DOI] [PubMed] [Google Scholar]
  • 128.Ishikawa K and Hayashi J (2010) A novel function of mtDNA: its involvement in metastasis. Ann. N. Y. Acad. Sci 1201, 40–43 [DOI] [PubMed] [Google Scholar]
  • 129.Imanishi H et al. (2011) Mitochondrial DNA mutations regulate metastasis of human breast cancer cells. PLoS One 6, e23401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Yuan Y et al. (2015) Nonsense and missense mutation of mitochondrial ND6 gene promotes cell migration and invasion in human lung adenocarcinoma. BMC Cancer 15, 346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Tang S et al. (2010) Left ventricular noncompaction is associated with mutations in the mitochondrial genome. Mitochondrion 10, 350–357 [DOI] [PubMed] [Google Scholar]
  • 132.Ji YC et al. (2011) The mitochondrial ND5 T12338C mutation may be associated with Leber’s hereditary optic neuropathy in two Chinese families. Yi Chuan 33, 322–328 (in Chinese) [DOI] [PubMed] [Google Scholar]
  • 133.Kenny TC et al. (2017) Selected mitochondrial DNA landscapes activate the SIRT3 axis of the UPR(mt) to promote metastasis. Oncogene 36, 4393–4404 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.LeBleu VS et al. (2014) PGC-1alpha mediates mitochondrial biogenesis and oxidative phosphorylation in cancer cells to promote metastasis. Nat. Cell Biol 16, 992–1003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Goh J et al. (2011) Mitochondrial targeted catalase suppresses invasive breast cancer in mice. BMC Cancer 11, 191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Fatemie S et al. (2012) Breast tumors in PyMT transgenic mice expressing mitochondrial catalase have decreased labeling for macrophages and endothelial cells. Pathobiol. Aging Age Relat. Dis 2, 17391. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 137.Bussard KM and Siracusa LD (2017) Understanding mitochondrial polymorphisms in cancer. Cancer Res. 77, 6051–6059 [DOI] [PubMed] [Google Scholar]
  • 138.Chattopadhyay M et al. (2022) The portrait of liver cancer is shaped by mitochondrial genetics. Cell Rep. 38, 110254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Chang X et al. (2022) Identification of mitochondrial DNA variants associated with risk of neuroblastoma. J. Natl. Cancer Inst 114, 910–913 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Zhou R et al. (2011) A role for mitochondria in NLRP3 inflammasome activation. Nature 469, 221–225 [DOI] [PubMed] [Google Scholar]
  • 141.Kelley N et al. (2019) The NLRP3 inflammasome: an overview of mechanisms of activation and regulation. Int. J. Mol. Sci 20, 3328. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Bufe B et al. (2015) Recognition of bacterial signal peptides by mammalian formyl peptide receptors: a new mechanism for sensing pathogens. J. Biol. Chem 290, 7369–7387 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Redza-Dutordoir M and Averill-Bates DA (2016) Activation of apoptosis signalling pathways by reactive oxygen species. Biochim. Biophys. Acta 1863, 2977–2992 [DOI] [PubMed] [Google Scholar]
  • 144.West AP et al. (2011) Mitochondria in innate immune responses. Nat. Rev. Immunol 11, 389–402 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Pavlides S et al. (2009) The reverse Warburg effect: aerobic glycolysis in cancer associated fibroblasts and the tumor stroma. Cell Cycle 8, 3984–4001 [DOI] [PubMed] [Google Scholar]
  • 146.Pavlides S et al. (2012) Warburg meets autophagy: cancer-associated fibroblasts accelerate tumor growth and metastasis via oxidative stress, mitophagy, and aerobic glycolysis. Antioxid. Redox Signal 16, 1264–1284 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Martinez-Outschoorn UE et al. (2010) Oxidative stress in cancer associated fibroblasts drives tumor-stroma co-evolution: a new paradigm for understanding tumor metabolism, the field effect and genomic instability in cancer cells. Cell Cycle 9, 3256–3276 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Pavlides S et al. (2010) Loss of stromal caveolin-1 leads to oxidative stress, mimics hypoxia and drives inflammation in the tumor microenvironment, conferring the ‘reverse Warburg effect’: a transcriptional informatics analysis with validation. Cell Cycle 9, 2201–2219 [DOI] [PubMed] [Google Scholar]
  • 149.Bonuccelli G et al. (2010) The reverse Warburg effect: glycolysis inhibitors prevent the tumor promoting effects of caveolin-1 deficient cancer associated fibroblasts. Cell Cycle 9, 1960–1971 [DOI] [PubMed] [Google Scholar]
  • 150.Narayanan S et al. (2020) Reprogramming of cancer cell Metabolism: Warburg and reverse Warburg hypothesis. In Cancer Cell Metabolism: A Potential Target for Cancer Therapy (Kumar D, ed.), pp. 15–26, Springer Nature [Google Scholar]
  • 151.Mitchell MI and Engelbrecht AM (2017) Metabolic hijacking: a survival strategy cancer cells exploit? Crit. Rev. Oncol. Hematol 109, 1–8 [DOI] [PubMed] [Google Scholar]
  • 152.DeWeerdt S (2015) Microbiome: microbial mystery. Nature 521, S10–S11 [DOI] [PubMed] [Google Scholar]
  • 153.Buck MD et al. (2016) Mitochondrial dynamics controls T cell fate through metabolic programming. Cell 166, 63–76 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Majmundar AJ et al. (2010) Hypoxia-inducible factors and the response to hypoxic stress. Mol. Cell 40, 294–309 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Chandel NS et al. (2000) Reactive oxygen species generated at mitochondrial complex III stabilize hypoxia-inducible factor-1alpha during hypoxia: a mechanism of O2 sensing. J. Biol. Chem 275, 25130–25138 [DOI] [PubMed] [Google Scholar]
  • 156.Young SD et al. (1988) Hypoxia induces DNA overreplication and enhances metastatic potential of murine tumor cells. Proc. Natl. Acad. Sci. U. S. A 85, 9533–9537 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Gilkes DM et al. (2014) Hypoxia and the extracellular matrix: drivers of tumour metastasis. Nat. Rev. Cancer 14, 430–439 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Kim JW et al. (2007) Effects of hypoxia on tumor metabolism. Cancer Metastasis Rev. 26, 291–298 [DOI] [PubMed] [Google Scholar]
  • 159.Semenza GL (2000) Hypoxia, clonal selection, and the role of HIF-1 in tumor progression. Crit. Rev. Biochem. Mol. Biol 35, 71–103 [DOI] [PubMed] [Google Scholar]
  • 160.Semenza GL (2000) HIF-1: using two hands to flip the angiogenic switch. Cancer Metastasis Rev. 19, 59–65 [DOI] [PubMed] [Google Scholar]
  • 161.Aran D et al. (2017) Comprehensive analysis of normal adjacent to tumor transcriptomes. Nat. Commun 8, 1077. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Magee R et al. (2019) TRNA-derived fragments as sex-dependent circulating candidate biomarkers for Parkinson’s disease. Parkinsonism Relat. Disord 65, 203–209 [DOI] [PubMed] [Google Scholar]
  • 163.Tang L (2022) Base editing in mitochondrial DNA. Nat. Methods 19, 640. [DOI] [PubMed] [Google Scholar]

RESOURCES