You are on page 1of 19

http://www-ermm.cbcu.cam.ac.

uk

expert reviews

in molecular medicine

Mitochondria as targets for detection and treatment of cancer


Josephine S. Modica-Napolitano and Keshav K. Singh
Mitochondria are dynamic intracellular organelles that play a central role in oxidative metabolism and apoptosis. The recent resurgence of interest in the study of mitochondria has been fuelled in large part by the recognition that genetic and/or metabolic alterations in this organelle are causative or contributing factors in a variety of human diseases including cancer. Several distinct differences between the mitochondria of normal cells and cancer cells have already been observed at the genetic, molecular and biochemical levels. As reviewed in this article, certain of these alterations in mitochondrial structure and function might prove clinically useful either as markers for the early detection of cancer or as unique molecular sites against which novel and selective chemotherapeutic agents might be targeted.
Early studies of differences between the mitochondria of normal cells and those of cancer cells focused on the respiratory deficiencies common to rapidly growing cancer cells. This led Otto Warburg to propose in 1930 that respiratory deficiency might result in dedifferentiation of cells and hence neoplastic transformation (Refs 1, 2). Other early studies suggested that transformation was the result of submolecular micro-electronic changes involving errant dismantling and rebuilding of the electron transport chain during cell division (Ref. 3), and that mutations in mitochondria might cause cancers (Ref. 4). More recently, a great deal of research has been conducted that substantiates, extends and provides new insight into the role of mitochondria in cancer (Refs 5, 6, 7, 8). This review summarises the important aspects of mitochondrial structure and function, highlights the observed differences in mitochondria between normal and cancer cells, and discusses how these differences might be exploited in the detection and treatment of cancer.

Mitochondria structure, function and genome Historical perspective


Early cytological studies (reviewed in Ref. 9) indicated the presence of subcellular granules similar in size and shape to bacteria in a variety of different cell types. In 1890, Altman postulated that these granules, which he termed bioblasts,

Josephine S. Modica-Napolitano Associate Professor, Department of Biology, Merrimack College, North Andover, MA 01845, USA. Tel: +1 978 837 5000 x4459; Fax: +1 978 837 5029; E-mail: josephine.modicanapolitano@merrimack.edu Keshav Singh (corresponding author) Assistant Professor, Johns Hopkins Oncology Center, Bunting-Blaustein Cancer Research Building, 1650 Orleans St, Baltimore, MD 21231, USA. Tel: +1 410 614 5128; Fax: +1 410 502 7234/7244; E-mail: singhke@jhmi.edu

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews

in molecular medicine

were the basic units of cellular activity. Interestingly, Altman further speculated that bioblasts were capable of an independent existence, yet formed a colonial association with the cytoplasm of a host cell, and that it was through this association that the host cell acquired the properties of life. The term mitochondrion, meaning thread-like granule, was first applied to these subcellular structures by Benda in 1898 . During the period 19001930, most cytologists recognised the mitochondrion as a well-defined and ubiquitous organelle, although at that time there was no agreement about its function. The identification of mitochondria as centres of energy metabolism came at the heels of refinements in cell fractionation techniques during the late 1940s, which allowed the successful separation of relatively pure, functionally intact mitochondria from other cellular components in liver cell homogenates (Refs 10, 11). By 1949, these mitochondrial fractions were shown to contain succinate oxidase and cytochrome oxidase activities, as well as the enzyme systems required for fatty acid oxidation and the citric acid cycle (Refs 12, 13; reviewed in Ref. 9). Today, it is known that mitochondria produce up to 80% of the energy needs of a cell and perform a host of additional cellular functions.

located between the outer and inner membranes; and the matrix is the space enclosed by the inner mitochondrial membrane. By contrast to the static, cigar-shaped organelles commonly observed in electron micrographs, living cells stained with the lipophilic cation rhodamine 123 (Rh123) and observed by fluorescence microscopy reveal mitochondria as a dynamic network of long filamentous structures, capable of profound changes in size, form and location (Ref. 14). These mitochondria can be seen extending, contracting, fragmenting and even fusing with one another as they move in three dimensions throughout the cytoplasm. Interestingly, the treatment of cells with microtubule-depolymerising agents has been shown to result in an altered distribution of mitochondria (Refs 15, 16). This suggests that mitochondria are associated with and travel along a molecular highway composed of a cytoplasmic microtubule network.

Mitochondrial function
Mitochondria play a central role in oxidative metabolism in eukaryotes (reviewed in Ref. 9). In the catabolism of carbohydrates (Fig. 1a), this begins with the transport of pyruvate from the cytosol into the mitochondrion, and its subsequent oxidative decarboxylation to acetyl CoA by a soluble, multi-enzyme pyruvate dehydrogenase complex, which is located in the mitochondrial matrix. The oxidation of acetyl CoA is achieved by a cyclic process involving eight catalytic steps. This process is known as either the citric acid or the tricarboxylic acid (TCA) cycle. All but one of the TCA cycle enzymes are soluble proteins found in the inner mitochondrial matrix compartment. The single insoluble enzyme, succinate dehydrogenase, is tightly bound to the matrix side of the inner mitochondrial membrane. Each round of the TCA cycle results in the production of two molecules of CO2, three molecules of reduced nicotinamide adenine dinucleotide (NADH), one molecule of reduced flavin adenine dinucleotide (FADH2), and one molecule of GTP (the energetic equivalent of ATP). The next stage of aerobic metabolism is oxidative phosphorylation, an energy-generating process that couples the oxidation of respiratory substrates (such as the NADH and FADH 2 generated through the TCA cycle) to the synthesis of ATP. Substrate oxidation involves a series of

Mitochondrial structure
The mitochondria of different tissues are similar in their gross morphology (reviewed in Ref. 9). In electron micrographs of fixed tissue specimens, mitochondria are most commonly observed as oval particles, 12 m in length and 0.51 m in width. These dimensions approximate to those of the bacterium Escherichia coli. The organelle is bounded by two membranes. The peripheral, or outer, membrane encloses the entire contents of the mitochondrion. The inner membrane has a much greater surface area and forms a series of folds or invaginations, called cristae, which project inward towards the interior space of the organelle. The total surface area of the inner membrane varies considerably depending upon the tissue and type of cell. Since the enzymes involved in oxidative phosphorylation are located on the inner mitochondrial membrane, its surface area and number of cristae are generally correlated with the degree of metabolic activity exhibited by a cell. The spatial arrangement of the outer and inner membranes creates two distinct internal compartments: the intermembrane space is

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews
b Fatty acid oxidation Fatty acids c Urea cycle Ornithine Urea

in molecular medicine

a Carbohydrate metabolism Pyruvate

Cytosol Membrane Soluble + enzymes in matrix -oxidation Citrulline Urea cycle

(Mostly) Acetyl CoA Succinate dehydrogenase (inner membrane) + Oxidation CO2 GTP NADH FADH2 Oxidative phosphorylation Respiratory enzyme + complexes I-V (inner membrane) ATP Adenine nucleotide + translocase Energy use: in cytosol (and by mitochondrion) TCA V cycle Acetyl CoA

(Sometimes) e.g. fasting or disease

Schematic of key mitochondrial metabolic pathways


Expert Reviews in Molecular Medicine C 2002 Cambridge University Press Figure 1. Schematic of key mitochondrial metabolic pathways. (a) Carbohydrate metabolism. Pyruvate produced from glycolysis undergoes oxidative decarboxylation to acetyl CoA, which is then oxidised in an eight-step process known as the tricarboxylic acid (TCA) cycle. The respiratory substrates NADH and FADH2 generated through the TCA cycle are next oxidised in a process coupled to ATP synthesis. Electrons are transferred from NADH and FADH2 to oxygen via enzyme complexes located on the inner mitochondrial membrane. Three of the electron carriers (complexes I, III and IV) are proton pumps, and couple the energy released by electron transfer to the translocation of protons from the matrix side to the external side of the inner mitochondrial membrane. Energy stored in the resulting proton gradient (i.e. the proton-motive force) is used to drive the synthesis of ATP via the mitochondrial enzyme ATP synthetase (complex V). (b) Fatty acid oxidation. Fatty acids undergo oxidative decarboxylation in the mitochondrial matrix to give acetyl CoA, which is fed into the TCA cycle, and new acyl CoA molecules that are successively shortened with each round of the cycle. Under certain conditions (e.g. fasting), acetyl CoA molecules are converted into ketones for use as an alternative energy source. (c) Urea cycle. Amino acid degradation resulting in excretion of nitrogen as urea occurs partly in the mitochondrion. The mitochondrion is also essential for several other processes (not shown), including gluconeogenesis, regulation of cytosolic NAD+, intracellular homeostasis of inorganic ions, and apoptosis (fig001ksb).
Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

Matrix of mitochondrion

Oxidative decarbox -ylase

Oxidative decarbox -ylase

Fatty acid V cycle

Matrix enzymes

Ketones

Energy source: brain heart, etc.

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews

in molecular medicine

respiratory enzyme complexes that are located on the inner mitochondrial membrane and are capable of accepting and donating electrons in a specific sequence based on their relative oxidationreduction potentials and substrate specificity. Complex I (NADH-ubiquinone reductase) transfers electrons from NADH to the mobile electron carrier ubiquinone, or coenzyme Q. It is the largest and most labile of all the respiratory enzyme complexes. In bovine heart, for example, complex I comprises at least 41 different protein subunits. Complex II (succinate-ubiquinone reductase) transfers reducing equivalents from succinate to ubiquinone. It comprises four protein subunits, one of which is the FADH 2-linked TCA cycle enzyme succinate dehydrogenase. Complex III (ubiquinone-cytochrome c reductase) is an 11subunit respiratory enzyme complex involved in the transfer of electrons from membrane-bound ubiquinone to oxidised cytochrome c, another mobile electron carrier located on the outer surface of the inner mitochondrial membrane. Complex IV, or cytochrome c oxidase (COX), is the terminal electron acceptor. It comprises 13 different protein subunits and functions in the transfer of electrons from reduced cytochrome c to molecular oxygen, to form H2O. The energy released by the exergonic transfer of electrons from respiratory substrate to oxygen is coupled to the translocation of protons from the matrix side to the external side of the inner mitochondrial membrane at three sites: respiratory enzyme complexes I, III and IV. In intact, well-coupled mitochondria, the inner membrane is relatively impermeable to the back flow of these protons. According to the Chemiosmotic Hypothesis, which was first proposed by Peter Mitchell in 1961 (and for which he received the Nobel Prize in Chemistry in 1978), the energy stored in the resulting proton gradient (i.e. the proton-motive force) is used to drive the synthesis of ATP via complex V, the mitochondrial enzyme ATP synthetase (Ref. 17). The ATP that is produced by aerobic metabolism and not used by the mitochondrion is transported across the inner mitochondrial membrane in exchange for cytosolic ADP by the enzyme adenine nucleotide translocase (ANT). This exchange ensures not only the availabilty of mitochondrial ADP, which is the principal control molecule for the rate of oxidative phosphorylation, but also the availability of

cytosolic ATP. Oxidative phosphorylation thus supplies a majority of the cellular energy produced under aerobic conditions and required to sustain cell viability and normal cell functions. Fatty acid oxidation is another important metabolic activity located in the mitochondria (Fig. 1b). The beta-oxidation pathway involves four separate enzymes that are soluble in the mitochondrial matrix and that function in a repetitive cycle. With each round of the cycle, a fatty acid undergoes oxidative decarboxylation to produce one molecule of acetyl CoA and one molecule of a new acyl CoA that is two carbons shorter than the starting fatty acid. The process continues until the original fatty acid molecule is completely degraded to acetyl CoA (for example, the 16-carbon palmitoyl CoA would undergo seven rounds of beta-oxidation to yield eight molecules of acetyl CoA). The acetyl CoA molecules thus generated normally enter into the TCA cycle where they undergo oxidation to CO2. However, during conditions of prolonged fasting and starvation, or in certain metabolic diseases (e.g. diabetes mellitus), the acetyl CoA molecules generated by fatty acid oxidation are converted into ketones (e.g. -hydroxybutyrate, acetoacetate and acetone) by enzymes also located in the mitochondrial matrix. These molecules are then transported through the blood to other tissues, such as brain and heart, where they are used as an alternative energy source to glucose. In addition to its central role in oxidative metabolism, the mitochondrion is involved in a variety of other important cellular functions. For example, certain enzymes of the urea cycle (Fig. 1c) and gluconeogenesis are located in the mitochondrial matrix. Mitochondria are involved in the regeneration of cytosolic NAD+ (required for the substrate-level phosphorylation step in glycolysis) and in the intracellular homeostasis of inorganic ions such as calcium and phosphate. A wealth of recent studies show that mitochondria also play an integral role in the cascade of intracellular events that lead to apoptosis, or programmed cell death (Refs 18, 19).

Mitochondrial genome
Mammalian cells typically contain 103104 copies of mitochondrial DNA (mtDNA). The genome is a 16.5 kb closed-circular, double-helical molecule that encodes two rRNAs, 22 tRNAs and 13 polypeptides (reviewed in Ref. 5). Each of these polypeptides is a highly hydrophobic subunit of

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews

in molecular medicine

one of four respiratory enzyme complexes localised to the inner mitochondrial membrane. They include seven subunits of respiratory enzyme complex I, one subunit of complex III, three subunits of complex IV, and two subunits of complex V. All other mitochondrial proteins, including those involved in the replication, transcription and translation of mtDNA, are encoded by nuclear genes and are targeted to the mitochondrion by a specific transport system (Ref. 20). In humans and other mammals, mitochondrial genes display a maternal inheritance (i.e. are inherited from the female parent). This is probably because the number of mtDNA copies in the egg is typically 103-fold greater than that in the sperm (Ref. 21). Alternatively, paternal genomes and organelles might be preferentially degraded in the zygote (Refs 22, 23). Although the mitochondrial and nuclear genomes are physically distinct, the high degree of functional interdependence between them is suggestive of a hostparasite relationship. The endosymbiont theory proposes that early in the evolution of the eukaryote, a primitive protoeukaryote cell that was incapable of aerobic respiration served as host to a eubacterium with the unique capacity for oxidative metabolism (Ref. 24). During the early stages of this endosymbiotic association, the eubacterium retained its genetic autonomy. In time, however, most of its genetic material was transferred to the nuclear genome of the host. The resulting mitochondrion retained only those few (i.e. 13) genes encoding polypeptides that are essential to aerobic ATP production yet have a hydrophobicity that precludes nuclear synthesis and cytoplasmic transport to mitochondria. It is of interest to recall Altmans perceptive characterisation of mitochondrial function, and his suggestion of a colonial association between the newly discovered bioblasts and the host cell within which they reside.

with Lebers hereditary optic neuropathy and familial mitochondrial encephalomyopathy (Refs 27, 28, 29). Since then, there has been a steady growth in the list of diseases associated with mitochondrial dysfunction arising from mtDNA mutations (Table 1). Although mtDNA represents less than 1% of total cellular DNA, its gene products are essential for normal cell function. Unlike nuclear DNA, mammalian mtDNA contains no introns, has no protective histones and is exposed to deleterious

Table 1. Mitochondrial diseasesa (tab001ksb)


Tissue/organ affected
Blood Brain

Clinical condition
Pearsons syndrome Seizures Myoclonus Ataxia Stroke Demetia Migraine Pseudo-obstruction Optic neuropathy Ophthalmoplegia Retinopathy Conduction disorder WolffParkinsonWhite syndrome Cardiomyopathy Sensorineural hearing loss Fanconis syndrome Glomerulopathy Hepatopathy Myopathy Neuropathy

Colon Eye

Heart

Inner ear Kidney Liver Skeletal muscle

Mitochondria and disease


The first mitochondrial disease was described by Rolf Luft in 1962 (Ref. 25), a year prior to the discovery of mtDNA (Ref. 26). At the genetic level, mitochondrial disease was not definitively described until 1988 with the identification of mtDNA mutations in patients with mitochondrial myopathies, as well as the molecular genetic characterisation of patients

Data in the table are derived from Ref. 38. Mitochondrial diseases can arise from mutations in nuclear DNA or in mitochondrial DNA. Deficits in ATP production might result from altered functions of proteins involved directly in oxidative phosphorylation or involved in communication between the nucleus and mitochondria, and might have deleterious effects on several organ systems. Many mitochondrial diseases are so new that they have not yet been mentioned in the medical textbooks or in the medical literature.

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews
are being described every year (Ref. 37). Thus, mitochondria play a central role in many diseases that can affect any organ, at any age.

in molecular medicine

reactive oxygen species generated by oxidative phosphorylation. In addition, replication of mtDNA might be error prone (Refs 8, 30, 31, 32, 33, 34). The accumulation of mutations in mtDNA is approximately tenfold greater than that in nuclear DNA (Refs 35, 36). Inherited disorders of mitochondria produce childhood and adult diseases with a variety of clinical symptoms (Ref. 37). Mitochondrial dysfunction has been found frequently as the basis of developmental defects. It is estimated that of the 4 million children born each year in the USA, up to 4000 develop diseases related to mitochondrial dysfunction (Refs 5, 30). Congenital mitochondrial diseases such as KearnsSayre/ chronic progressive external opthalmoplegia and Lebers hereditary optic neuropathy are either maternally inherited or are derived from a founder mutation early in embryogenesis. Since mitochondria perform a variety of different functions in different tissues, since the proportion of normal to mutated mtDNA can vary, and since tissues have different aerobic dependencies, each mtDNA mutation can produce a wide spectrum of phenotypes that have proven to be extremely perplexing to scientists and physicians. Mitochondrial dysfunction is also increasingly recognised as an important cause of adult human pathology (reviewed in Ref. 5). Dysfunctional mitochondria are found in diverse adult-onset diseases, including diabetes, cardiomyopathy, infertility, migraine, blindness, deafness, kidney and liver diseases, and stroke. The accumulation of somatic mutations in mtDNA has been suggested to play a causative or contributing role in aging, in age-related neurodegenerative disorders such as Parkinsons, Alzheimers and Huntingtons disease, and in cancer (see below). Other adult-onset pathologies might result from the biochemical toxicity or mtDNA damage caused by various drugs, including those used against human immunodeficiency virus (HIV) (Ref. 38), or by other endogenous or environmental agents (Refs 30, 39). Although human mitochondrial diseases are often multisystem disorders, constitutively highly oxidative tissues such as myocardium, brain and kidney, as well as episodically oxidative tissues such as skeletal muscle, are especially vulnerable to mtDNA damage (Refs 35, 36). To date, over 100 point mutations and 200 deletions and rearrangements have been shown to be associated with mitochondrial disease and new mutations

Mitochondria and cancer Phenotypic differences in tumour mitochondria


Cancer cells have an altered metabolism that includes: a higher rate of glycolysis (Ref. 39), an increased rate of glucose transport (Ref. 40), increased gluconeogenesis (Ref. 41), reduced pyruvate oxidation and increased lactic acid production (Ref. 42), increased glutaminolytic activity (Ref. 43), reduced fatty acid oxidation (Ref. 44), increased glycerol and fatty acid turnover (Ref. 45), modified amino acid metabolism (Ref. 46), and increased pentose phosphate pathway activity (Ref. 47). Mitochondria are involved either directly or indirectly in many aspects of altered metabolism in cancer cells (Ref. 48) and several notable differences between the mitochondria of normal versus transformed cells have been discovered (reviewed in Refs 49, 50 and 51). For example, various tumour cell lines exhibit differences in the number, size and shape of their mitochondria relative to normal controls. The mitochondria of rapidly growing tumours tend to be fewer in number, smaller and have fewer cristae than mitochondria from slowly growing tumours; the latter are larger and have characteristics more closely resembling those of normal cells. Interestingly, the usually benign oncocytoma of thyroid, salivary gland, kidney, parathyroid and breast is characterised by the presence of cells containing abnormally large numbers of mitochondria, and high levels of oxidative enzymes (Ref. 52). The ultrastructural features of mitochondria in these cells show similarities with mitochondrial encephalomyopathies, where mitochondria are found as large aggregates and display a variety of morphological alterations. MtDNA mutations are also commonly found in oncocytic tumours (Ref. 52). Alterations in the molecular composition of the inner membranes of tumour mitochondria have also been noted (Refs 53, 54, 55, 56, 57, 58). Polypeptide profiles of normal liver versus hepatoma mitochondria demonstrate differences in the appearance and/or relative abundance of several protein subunits. One major band that is deficient or absent in several tumours studied has a mobility near or equal to the B subunit of the F1-

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews

in molecular medicine

ATPase (approximately 57 kDa). Other bands that are present in tumour mitochondria appear to be deficient or absent in control mitochondria. In addition, analysis of the inner membrane lipid composition of various tumour mitochondria has indicated elevated levels of cholesterol, varying total phospholipid content, and/or changes in the amount of individual phospholipids relative to normal controls. Many differences in the mitochondria of normal versus transformed cells have also been noted with regard to: (1) the preference for substrates oxidised; (2) the magnitude of the acceptor control ratio; (3) the rates of electron and anion transport; (4) the capacity to accumulate and retain calcium; (5) the amounts and forms of DNA; and (6) the rates of protein synthesis and organelle turnover. However, there is apparently no universal metabolic alteration that is common to all tumours. For example, although the pathogenesis of prostate cancer involves the mitochondrial metabolic transformation of citrateproducing cells to citrate-oxidising cells, this metabolic abnormality is not reported in other cancers (Ref. 59). Additionally, it is important to note that the altered metabolism in cancer cells is probably not the cause of malignancy but, rather, a secondary, albeit essential, adaptation to support malignant activities (Ref. 59).

Mitochondria of neuron revealed by staining with a rhodamine 123 derivative


Expert Reviews in Molecular Medicine C 2002 Cambridge University Press Figure 2. Mitochondria of neuron revealed by staining with a rhodamine 123 derivative. Neurons were stained with tetramethylrhodamine, ethyl ester, perchlorate (TMRE). TMRE accumulates in mitochondria and emits high fluorescence. The cell images were recorded with a digital camera and processed to generate high-resolution images. The image is courtesy of Dr Gary E. Gibson, Weill Medical College of Cornell University, Burke Medical Research Institute, White Plains, NY 10605, USA. Magnification, ~X760 (fig002ksb).

Rh123 uptake
In the early 1980s, Chen and colleagues discovered an interesting phenotype that was found to be common to nearly all types of carcinoma tested (Refs 14, 60, 61). It was observed that the lipophilic cation Rh123 could serve as a highly specific vital stain for mitochondria, providing lowbackground, high-resolution fluorescent images of the organelle in a variety of cell types (Fig. 2). It was further observed that relative to the mitochondria of normal epithelial cells, the mitochondria of carcinoma cells displayed an increased uptake and prolonged retention of Rh123, and that this phenomenon correlated with a selective cytotoxicity for carcinoma cells in vitro and in vivo (Refs 62, 63, 64). For example, whereas Rh123 was shown to have a minimal effect on the clonal growth of those cell types that display little uptake and short retention of the dye (e.g. primary cultures of normal mouse bladder epithelial cells; and CCL-34 and BSC-1, the non-tumourigenic dog and monkey kidney epithelial cell lines, respectively), it markedly inhibited the clonal

growth of those cultured carcinomas cell lines that display high uptake and prolonged retention of Rh123 (e.g. MB49, the transformed mouse epithelial cell line; and MCF-7 and HUT, the human breast and lung carcinoma cell lines, respectively). In addition, at a constant exposure of 10 g/ml Rh123, greater than 50% cell death occurred within seven days in 9/9 of the carcinoma cell types tested, whereas 6/6 control epithelial cell types remained unaffected. Standard chemotherapeutic agents such as arabinosyl cytosine (Ara-c) and methotrexate exhibit no such selectivity for carcinoma cells. In vivo, Rh123 was shown to prolong the survival of mice implanted with Ehrlich ascites tumour or MB49 mouse bladder carcinoma cells as much as 260%, although the extent of survival prolongation was highly dependent on the dose and schedule of administration of the dye. As expected, Rh123 did not significantly prolong the survival rate of mice implanted with tumours of cell types shown to be short retainers of the dye (e.g. L1210 and P388 leukaemias, and B6 melanoma).

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews

in molecular medicine

Increased membrane potential in carcinoma cells


It became apparent from early studies that two chemical properties of Rh123 were important in promoting its uptake into mitochondria. The first is the lipophilicity of Rh123, which allows the compound to penetrate the hydrophobic barriers of the plasma and mitochondrial membranes; and the second is its electrical charge rhodamines that are positively charged at physiological pH stain mitochondria specifically, whereas the neutral rhodamines do not (Ref. 14). Initially, there was much indirect evidence to suggest that Rh123 uptake occurs as a function of the magnitude of the mitochondrial membrane potential. For example, the addition of ionophores that dissipate the mitochondrial electrical gradient (such as valinomycin or dinitrophenol), or respiratory inhibitors that prevent the establishment of the electrical gradient (such as cyanide, antimycin or rotenone), were demonstrated to diminish mitochondrial-specific fluorescence in cells prestained with the dye (Ref. 61). Later, experimental manipulation of the membrane potential in isolated mitochondria and concurrent measurement of the amount of Rh123 associated with the organelle definitively established that Rh123 is concentrated by cells and into mitochondria in response to negative-inside transmembrane potentials (Ref. 65). Furthermore, it was determined that the mitochondrial membrane potential of carcinoma cells is approximately 60 mV higher than that of control epithelial cells (Ref. 65). Since Rh123 distributes across the inner mitochondrial membrane in accordance with the Nernst equation (Ref. 66), this difference alone is sufficient to account for a tenfold greater accumulation of the compound in carcinoma versus control epithelial mitochondria. In whole cells, however, the plasma membrane potential pre-concentrates Rh123 relative to the external medium, thus affecting the cytoplasmic concentration of Rh123 and the amount of dye available for mitochondrial uptake. The higher plasma membrane potential observed in some carcinoma cells versus control epithelial cell types therefore further contributes to increased Rh123 accumulation in carcinoma mitochondria (Ref. 67). Finally, Rh123 was found to exhibit a concentration-dependent toxicity in mitochondria by inhibition of ATP synthetase (Refs 68, 69). Since mitochondria are the primary sites of ATP synthesis in cells undergoing aerobic metabolism,

selective mitochondrial toxicity in carcinoma cells resulting from enhanced uptake and retention of Rh123 provided the basis for the selective anticarcinoma activity displayed by this compound.

Delocalised lipophilic cations


Rh123 provided a prototype for a new class of anti-cancer agents that exploit the difference in mitochondrial membrane potential between normal epithelial and carcinoma cells to achieve a selective mitochondrial toxicity and consequent selective cytotoxicity for carcinoma cells. In the past few years, several members of this class of compounds, known collectively as delocalised lipophilic cations (DLCs), have exhibited at least some degree of efficacy in carcinoma cell killing in vitro and/or in vivo (Refs 70, 71, 72, 73, 74, 75). For example, dequalinium chloride (DECA) has demonstrated 100-fold greater inhibition of the clonal growth of carcinoma versus control epithelial cells in culture, anti-carcinoma activity in human colon adenocarcinoma cells injected subcutaneously in nude mice, and significant regression of tumours in rats carrying in situ mammary adenocarcinomas induced by dimethyl-bezanthracene (Refs 71, 72). The thiopyrylium AA-1 was shown to prolong the survival of mice implanted with either mouse bladder carcinoma, human melanoma, and human ovarian carcinoma cell lines, achieving treated:control ratios as high as 450% (Ref. 74). The rhodacyanine MKT-077 also appears particularly promising (Refs 73, 76, 77). In recent studies, MKT-077 demonstrated significant growth-inhibitory activity against a variety of keratin-positive human cancer cell lines, as measured by clonogenic assays and growth inhibition of cultured cells. In vivo, MKT-077 demonstrated significant anti-tumour activity in nude mice implanted with either the human melanoma LOX, the human renal carcinoma A498, or the human prostate carcinoma DU145, all of which are highly refractory to a variety of traditional therapies. As the first DLC with a favourable pharmacological and toxicological profile in preclinical studies, MKT-077 has undergone Phase I clinical trials, approved by the US Food and Drug Administration, for the treatment of carcinoma.

Mechanism of action of DLCs


It is of interest to note that, although all DLCs are taken up into mitochondria by a common

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews

in molecular medicine

mechanism (i.e. in response to negative-inside transmembrane potentials), the mechanism of mitochondrial toxicity exhibited by these compounds is quite varied. For example, among the DLCs that display a concentration-dependent toxicity to mitochondria, Rh123 and AA-1 inhibit mitochondrial ATP synthesis at the level of F0F1ATPase activity (Refs 68, 74), whereas DECA and certain DLC thiacarbocyanines interfere with NADH-ubiquinone reductase activity (Refs 70, 72). In addition, the selective cytotoxicity to carcinoma cells exhibited by MKT-077 in vitro and in vivo has been attributed to a selective inhibition of mitochondrial respiration in cancer cells, most probably as a result of a general perturbation of mitochondrial membranes and consequent inhibition of the activity of membrane-bound enzymes (Ref. 76). The selective cytotoxicity might also be a consequence of a mild to moderate degradative effect on mtDNA, but not nuclear DNA, of various carcinoma cell types (Ref. 76). It is of further interest to note that, although increased membrane potential is necessary to achieve selective cytotoxicity by DLCs, it alone is not sufficient. If this were the case then cardiac muscle cells, which have also been shown to exhibit a high mitochondrial membrane potential (Ref. 78), would be susceptible to the cytotoxic effects of these compounds. Yet significant cardiac toxicity has not been observed following in vivo administration of either MKT-077 or DECA. This suggests the sensitivity of any particular cell type to the effects of DLCs might depend on different cytoplasmic characteristics, such as those involving the kinetics of uptake and retention of the compound, or on properties inherent to the mitochondria, such as differential sensitivity of the target molecule against which the DLC exerts its cytotoxic effect.

DLCs and photochemotherapy


Some research groups have explored the use of certain DLCs in photochemotherapy (PCT), an investigational cancer treatment involving light activation of a photoreactive drug, or photosensitiser, that is selectively taken up or retained by malignant cells (Refs 79, 80, 81, 82). There has been considerable interest recently in PCT as a form of treatment for neoplasms of the skin, lung, breast, bladder, brain or any other tissue accessible to light transmitted either through the body surface or internally via fibre

optic endoscopes. Cationic photosensitisers are particularly promising as potential PCT agents. Like other DLCs, these compounds are concentrated by cells into mitochondria in response to negative-inside transmembrane potentials, and are thus selectively accumulated in the mitochondria of carcinoma cells. In response to localised photoirradiation, the photosensitiser can be converted to a more reactive and highly toxic species, thus enhancing the selective toxicity to carcinoma cells and providing a means of highly specific tumour cell killing without injury to normal cells. Several cationic photosensitisers have shown promise for use in PCT. For example, selective phototoxicity of carcinomas in vitro and in vivo has been observed for a series of triarylmethane derivatives (Ref. 83), and for 2-ethyl-1,3-dioxylene kryptocyanine (EDKC) (Ref. 77). Both Rh123 and the chalcogenapyrylium dye 8b have been evaluated as photosensitisers for the photochemotherapy of malignant gliomas (Refs 84, 85, 86). As is the case for the nonphotosensitising DLCs, the mitochondrion has been implicated as an important, perhaps primary, subcellular site of damage by these and several other cationic photosensitisers (Refs 79, 87, 88, 89, 90, 91). Again, the mechanisms of mitochondrial toxicity exhibited by these compounds have been shown to vary from an inhibition of NADHubiquinone reductase (e.g. in the case of EDKC, and the triarylmethane derivative VB-BO) to a non-specific perturbation of mitochondrial function most probably resulting from membrane damage induced by singlet oxygen (e.g. in the case of certain chalcogenapyrylium dyes). More recently, photoactivation has been shown to enhance the mitochondrial toxicity of MKT-077, with evidence for the involvement of lipid peroxidation in this process. These results have positive implications for the use of MKT-077 in PCT.

Future directions for research involving DLCs


Although the use of DLCs as anti-cancer agents has shown promise, as yet there is no real understanding of the biochemical basis for the increased mitochondrial membrane potential in carcinoma cells. Consequently, the choice for the design or selection of potentially therapeutic lipophilic cations has been based almost solely on physical properties (i.e. lipid solubilty,

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews

in molecular medicine

delocalisation of positive charge, etc.), and preliminary screening for the selective cytotoxicity of these compounds has been empirical. Although there is sufficient evidence to support the idea that a therapeutic mechanism based on differences in mitochondrial membrane potential might be exploited for the treatment of cancers, a knowledge of the specific biochemical alterations that account for increased mitochondrial membrane potential in these cells would undoubtedly lead to a more rational approach to the choice of highly selective DLCs for clinical use. To this end, a comprehensive, comparative biochemical analysis of mitochondria isolated from several control and carcinoma cell lines that display differences in mitochondrial membrane potential is currently under way (J.S. ModicaNapolitano, unpublished). This type of study will contribute to an understanding of the phenomenon of increased uptake of DLCs by carcinoma mitochondria and might also reveal molecular differences between the mitochondria of normal epithelial and carcinoma cells against which novel, selective and site-specific DLCs could be targeted. The information obtained will be used in the rational design of a more efficacious form of DLC that exhibits a dual selectivity for carcinoma cells based on both differential accumulation and selective action against a unique molecular target.

growth rates and degrees of differentiation display a decreased capacity for uncouplerstimulated ATP hydrolysis relative to that found in normal liver (Ref. 92). In addition, mitochondria isolated from biopsies of human hepatocellular carcinoma have decreased rates of respiration-linked ATP synthesis and a reduced phosphorylative capacity compared with normal human liver (Refs 93, 94). Furthermore, the measured maximal velocity for ATPase activity in submitochondrial particles isolated from hepatocellular carcinoma is considerably lower than that in normal liver. It has been suggested that these alterations in enzyme function might be associated with a decrease in immunodetectable levels of the B subunit of the F1 component of mitochondrial ATPase and/or with overexpression of the ATPase inhibitor protein (IF1) in tumour mitochondria (Refs 93, 94, 95).

COX activity
A comparison of COX activities in cultured carcinoma versus normal epithelial cells suggests a possible correlation between membrane potential and the activity of this enzyme. Measurements of COX activity in samples from the total cellular homogenate and the mitochondrial subfraction from the cultured human carcinoma cell lines MCF-7 (breast), T47D (breast) and DU-145 (prostate) demonstrate significantly lower specific activities of the enzyme compared with that measured in the normal monkey kidney epithelial cell line CV-1 (Ref. 96). Similar decreases in COX activity were found when comparing the specific activity of the enzyme in biopsies of human colonic adenocarcinoma versus normal colon mucosa (Ref. 97), and in cultured rat HC252 hepatoma cells versus non-neoplastic liver (Ref. 98). It is not clear whether the decrease in specific activity of COX in cancer cells can be explained by alterations in the level of gene expression. For example, in one study involving human colonic biopsies, the mean level of expression of mitochondrially encoded COX subunit III was found to be lower in carcinoma versus normal mucosa samples (Ref. 99). Cultured HT29 colon carcinoma cells also exhibited low levels of the COX III transcript; however, expression of COX III returned to higher (normal) levels when the cells were induced to differentiate by exposure to sodium butyrate. By contrast, increased

Molecular basis for increased membrane potential in carcinoma cells


It is logical to assume that the observed differences in the magnitude of the mitochondrial membrane potential between normal epithelial and carcinoma cells might arise from differences in the structure and function of one or more of those organelle components that serve to create and/or maintain the electrical gradient. Possibilities include differences in mitochondrial respiratory enzyme complexes, electron carriers, ATP synthetase, ANT and membrane lipid structure. Differences that affect electron transfer activity, or proton translocation, utilisation or conductance, are also candidates for the molecular basis for increased mitochondrial membrane potential in carcinoma cells. Interestingly, some differences of these types are already known to exist between the mitochondria of normal and malignant cells. For example, under certain assay conditions, mitochondria isolated from hepatomas of varying

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

10

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews

in molecular medicine

levels of RNA transcripts of the nuclear-encoded subunit COX IV and mitochondrially encoded COX subunits I and II have been observed in Zajdela hepatoma as compared with normal liver (Ref. 100). Alternatively, and perhaps more interestingly, the decreased specific activity of COX in tumour mitochondria might reflect a kinetic difference caused by genetic mutations that alter the structure and function of the enzyme. These same structural changes might be hypothesised to affect the proton-pumping capacity of the enzyme, and thus the magnitude of the mitochondrial membrane potential.

Mitochondrial permeability transition pore activity


The magnitude of the mitochondrial membrane potential might also be dependent upon membrane permeability as regulated generally by the membrane lipidprotein structure, or more specifically by the mitochondrial permeability transition pore (MPTP). This is a multi-protein structure formed at contact sites between the inner and outer mitochondrial membranes. It is a voltage-dependent, cyclosporin-A-sensitive, high-conductance inner membrane channel, the opening of which transiently depolarises the mitochondrial membrane potential. Although the complete physiological role of the MPTP is unknown at this time, it is probably involved in the early apoptotic changes that affect mitochondrial membrane permeability. Several differences have been found when comparing the MPTP in normal versus malignant cells. One of the key structural components of the MPTP complex is ANT. The primary role of ANT is to facilitate the one-for-one exchange of ATP (out) for ADP (in) across the inner mitochondrial membrane. More recently, it has been suggested that ANT also acts as a non-specific pore that renders the mitochondrial inner membrane permeable to solutes less than 1.5 kDa in size (Ref. 101). Interestingly, the adenine nucleotide exchange function of ANT is known to be decreased in certain hepatoma versus normal liver mitochondria (Refs 102, 103, 104). In addition, the sensitivity of this enzyme to bongkrekic acid, an inhibitor of both adenine nucleotide exchange and formation of the MPTP is also decreased in hepatoma versus normal liver (Refs 104, 105, 106). Furthermore, high transcript levels for ANT2, the gene encoding one of three isoforms of the

translocase, have been observed in several dedifferentiated, proliferating, renal tumour cell types, whereas expression of ANT2 is usually repressed in quiescent cells (Refs 107, 108). Additional known or putative MPTP components also exhibit alterations in gene expression between normal and cancer cells. Among those genes overexpressed in cancer cells are the anti-apoptotic oncogenes encoding Bcl-2 and Bcl-XL, which have a direct inhibitory effect on pore opening, and genes encoding the peripheral benzodiazepin receptor (PBR), the PBR-associated protein Prax-1, and mitochondrial creatine kinase (Refs 109, 110, 111, 112, 113, 114, 115, 116). Conversely, the expression of BAX, a proapoptotic, inner mitochondrial membrane protein that facilitates pore opening, has been shown to be reduced in some cancer cell lines (Refs 117, 118). However, whether these changes in gene expression contribute to steady-state differences in membrane permeability between normal epithelial and carcinoma cells has yet to be determined.

Mitochondrial DNA mutations in carcinoma cells


Mitochondrial dysfunction is one of the most profound features of cancer cells. Consistently, mutations in mtDNA have been reported in a variety of cancers. These include ovarian, thyroid, salivary, kidney, liver, lung, colon, gastric, brain, bladder, head and neck, and breast cancers, and leukaemia (see Table 2 for references). The types of mutations observed in mtDNA range from point mutations, to deletions and duplications. Most tumours contain homoplasmic (100% pure) mutant mtDNA because of the clonal nature of cancers (Refs 5, 6, 119). The abundance and homoplasmic nature of mitochondria make mtDNA an attractive molecular marker of cancer. Indeed, mutant mtDNA in tumour cells is reported to be 220 times as abundant as a mutated nuclear marker (Ref. 119). Furthermore, a recent study of mtDNA from patients with bladder, head and neck, and lung cancers reported that mutated mtDNA is readily detectable in urine, blood and saliva samples from these patients (Ref. 119). Thus, mtDNA mutation might prove to be an extremely useful biomarker for the detection of many cancers. Ongoing research on DNA repair genes involved in maintaining the genetic integrity of the mitochondrial genome combined with further

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

11

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews

in molecular medicine

Table 2. Mitochondrial DNA mutation in cancersa (tab002ksb)


Cancer
Leukaemia Ovary Thyroid Salivary Kidney Liver Lung Colon Gastric Brain Goiter Breast
a

Refs
123, 124, 125, 126, 127 128 129, 130 131, 132 133, 134, 135 136, 137 138 6, 139 140 141 52 142

Point mutations, deletions or duplications of mitochondrial DNA mutations are found in a wide range of cancers.

analysis of the nature of mtDNA mutations will greatly aid progress in this area. Mitochondrial dysfunction also has important implications in cancer therapy. This has been demonstrated by measuring the cell survival of a cervical tumour cell line (with parental mitochondrial function, i.e. Rho + ) and its derivative isogenic cell line that completely lacked mtDNA (with dysfunctional mitochondria, i.e. Rho0) after exposure to a variety of anti-cancer agents (Ref. 119). It was found that mitochondrial dysfunction leads to increased cell survival after exposure to cancer therapeutic agents such as adriamycin and porphyrin-catalysed phototoxicity. By contrast, no measurable difference was found in the cell survival of the Rho+ and Rho0 cells to high doses of ionising radiation. These results underscore the importance of the mitochondrial genome in development of cancer therapeutic drugs.

utilises lipophilic cations that accumulate selectively in carcinoma cells in response to increased mitochondrial membrane potential. An alternative strategy employs mitochondrial protein-import machinery to deliver macromolecules to mitochondria. For example, a mitochondrial signal sequence has been used to direct green fluorescent protein to mitochondria, which allows the visualisation of mitochondria within living cells (Ref. 121). Interestingly, certain short peptides readily penetrate the mitochondrial membrane and become toxic when internalised into the targeted cells by disruption of mitochondrial membranes (Ref. 122). Another chemotherapeutic strategy employs specific interaction of drugs with certain mitochondrial proteins (Ref. 101). Traditional chemotherapies, aimed at DNA replication in actively dividing cells, have achieved only limited success in the treatment of cancer largely because of their lack of specificity for cells of tumourigenic origin. It is important, therefore, to search for novel cellular targets that are sufficiently different between normal cells and cancer cells so as to provide a basis for selective cytotoxicity. As this review suggests, the mitochondrion is one such target.

Acknowledgements and funding


We thank the members of our laboratories for their contributions to this article. Our research has been supported by grants from the National Institutes of Health (RO1-097714, P50 CA88843, P20 CA86346), an American Heart Association Scientist Development Award (9939223N) to K.K.S., and a National Institutes of Health grant (R15 Ca78323-01S1) to J.S.M-N. We also thank Dr June R. Aprille, Tufts University, Medford, MA, USA and Lene J. Rasmussen, Roskilde University, Roskilde, Denmark for their helpful critique of this article.

Concluding remarks and clinical implications


The many distinct differences in mitochondrial structure and function between normal cells and cancer cells offer a unique potential for the clinical use of mitochondria as markers for the early detection of cancer. Additionally, these differences offer the possibility for the design and synthesis of effective anti-cancer agents that deliver potent mitochondrial inhibitors to selectively kill tumour cells (reviewed in Ref. 120). As summarised previously, one current chemotherapeutic strategy

References
1 Warburg, O. (1930) Metabolism of Tumors, Arnold Constable, London, UK 2 Warburg, O. (1956) On the origin of cancer cells. Science 123, 309-314 3 Szent-Gyorgyi, A. (1977) Electronic biology and cancer. In Search and Discovery: A Tribute to Albert Szent-Gyorgyi (Kaminer, B., ed.,), pp. 329335, Academic Press, New York, USA 4 Woods, M.W. and DuBuy, H.G. (1945) Cytoplasmic diseases and cancer. Science 102,

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

12

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews
Mutations in Aging, Disease and Cancer (Singh, K.K., ed.), pp. 147-165, Springer-Verlag, Berlin, Germany Zamzami, N. et al. (1996) Mitochondrial control of nuclear apoptosis. J Exp Med 183, 1533-1544, PubMed ID: 96261655 Schatz, G. (1996) The protein import system of mitochondria. J Biol Chem 271, 31763-31766, PubMed ID: 97112957 Birky, C.W.J. (1994) Relaxed and stringent genomes: why cytoplasmic genes dont obey Mendels laws. J Hered 85, 355-36624 Hiraoka, J. and Hirao, Y. (1988) Fate of sperm tail components after incorporation into the hamster egg. Gamete Res 19, 369-380, PubMed ID: 89065541 Kaneda, H. et al. (1995) Elimination of paternal mitochondrial DNA in intraspecific crosses during early mouse embryogenesis. Proc Natl Acad Sci U S A 92, 4542-4546, PubMed ID: 95273399 Gray, M.W. (1992) The endosymbiont hypothesis revisited. Int Rev Cytol 141, 233-357, PubMed ID: 93084476 Luft, R. et al. (1962) A case of severe hypermetabolism of nonthyroid origin with a defect in the maintenance of mitochondrial respiratory control: a correlated clinical, biochemical and morphological study. J Clin Invest 41, 1776-1804 Nass, S. and Nass, M.M.K. (1963) Intramitochondrial fibers with DNA characteristics. J Cell Biol 19, 613-629 Wallace, D.C. et al. (1988) Mitochondrial DNA mutation associated with Lebers hereditary optic neuropathy. Science 242, 1427-1430, PubMed ID: 89072713 Holt, I.J., Harding, A.E. and Morgan-Hughes, J.A. (1988) Deletions of muscle mitochondrial DNA in patients with mitochondrial myopathies. Nature 331, 717-719, PubMed ID: 88143157 Zeviani, M. et al. (1988) Deletions of mitochondrial DNA in Kearns-Sayre syndrome. Neurology 38, 1339-1346, PubMed ID: 88319290 Singh, K.K. et al. (2001) Inactivation of Saccharomyces cerevisiae OGG1 DNA repair gene leads to an increased frequency of mitochondrial mutants. Nucleic Acids Res 29, 1381-1388, PubMed ID: 21138442 Kunkel, T.A. and Loeb, L.A. (1981) Fidelity of mammalian DNA polymerases. Science 213, 765767, PubMed ID: 81249139 Matsukage, A., Bohn, E.W. and Wilson, S.H.

in molecular medicine

591-593 5 Singh, K.K. (1998) Mitochondrial DNA Mutations in Aging, Disease, and Cancer, Springer, New York, USA 6 Polyak, K. et al. (1998) Somatic mutations of the mitochondrial genome in human colorectal tumours. Nat Genet 20, 291-293, PubMed ID: 99021388 7 Kroemer, G. and Reed, J.C. (2000) Mitochondrial control of cell death. Nat Med 6, 513-519, PubMed ID: 20264559 8 Singh, K.K. (2000) Mitochondrion me and the mitochondrial journal. Mitochondrion 1, 1-2 9 Tzagoloff, A. (1982) Mitochondria, Plenum Press, New York, NY, USA 10 Claude, A. (1946) Fractionation of mammalian liver cells by differential centrifugation: II. Experimental procedures and results. J. Exp. Med. 84, 61 11 Hogeboom, G.H., Schneider, W.C. and Palade, G.E. (1948) Cytochemical studies of mammalian tissue. I. Isolation of intact mitochondria from rat liver; some biochemical properties of mitochondria and submiscroscopic particulate material. J. Biol. Chem. 172, 619-635 12 Green, D.E., Loomis, W.F. and Auerbach, V.H. (1948) Studies on the cyclophorase system. I. The complete oxidation of pyruvic acid to carbon dioxide and water. J. Biol. Chem. 172, 389-403 13 Kennedy, E.P. and Lehninger, A.L. (1949) Oxidation of fatty acids and tricarboxylic acid cycle intermediates by isolated liver mitochondria. J. Biol. Chem. 179, 957-972 14 Johnson, L.V., Walsh, M.L. and Chen, L.B. (1980) Localization of mitochondria in living cells with rhodamine 123. Proc Natl Acad Sci U S A 77, 990994, PubMed ID: 80145797 15 Aufderheide, K.J. (1980) Mitochondrial associations with specific microtubular components of the cortex of Tetrahymena thermophila. II. Response of the mitochondrial pattern to changes in the microtubule pattern. J Cell Sci 42, 247-260, PubMed ID: 80249708 16 Summerhayes, I.C., Wong, D. and Chen, L.B. (1983) Effect of microtubules and intermediate filaments on mitochondrial distribution. J Cell Sci 61, 87-105, PubMed ID: 83291373 17 Mitchell, P. (1961) Coupling of phosphorylation to electron and hydorgen transfer by a chemiosmotic type of mechanism. Nature 191, 144-148 18 Petit, P.X. and Kroemer, G. (1998) Mitochondrial regulation of apoptosis. In Mitochondial DNA

19

20

21

22

23

24

25

26

27

28

29

30

31

32

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

13

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews
glycerol kinetics in septic patients and in patients with gastrointestinal cancer. The response to glucose infusion and parenteral feeding. Ann Surg 205, 368-376, PubMed ID: 87183717 Souba, W.W. (1993) Glutamine and cancer. Ann Surg 218, 715-728, PubMed ID: 94079452 Boros, L.G. et al. (1998) Inhibition of the oxidative and nonoxidative pentose phosphate pathways by somatostatin: a possible mechanism of antitumor action. Med Hypotheses 50, 501506, PubMed ID: 98374098 Peluso, G. et al. (2000) Cancer and anticancer therapy-induced modifications on metabolism mediated by carnitine system. J Cell Physiol 182, 339-350, PubMed ID: 20119343 Pedersen, P.L. (1978) Tumor mitochondria and the bioenergetics of cancer cells. Prog Exp Tumor Res 22, 190-274, PubMed ID: 78227271 Weinhouse, S. (1955) Oxidative metabolism of neoplastic tissues. Adv. Cancer Res. 3, 269-325 Carafoli, E. (1980) Mitochondria and disease. Mol. Aspects Med. 3, 295-429 Maximo, V. and Sobrinho-Simoes, M. (2000) Hurthle cell tumours of the thyroid. A review with emphasis on mitochondrial abnormalities with clinical relevance. Virchows Arch 437, 107115, PubMed ID: 20445365 Chang, L.O., Schnaitman, C.A. and Morris, H.P. (1971) Comparison of the mitochondrial membrane proteins in rat liver and hepatomas. Cancer Res 31, 108-113, PubMed ID: 71125206 Irwin, C.C. and Malkin, L.I. (1976) Differences in total mitochondrial proteins and mitochondrially-synthesized proteins from rat liver and Morris hepatomas. Fed. Proc. Am. Soc. Exp. Biol. 35, 1583 Catterall, W.A. and Pedersen, P.L. (1971) Adenosine triphosphatase from rat liver mitochondria. I. Purification, homogeneity, and physical properties. J Biol Chem 246, 4987-4994, PubMed ID: 71288578 Catterall, W.A., Coty, W.A. and Pedersen, P.L. (1973) Adenosine triphosphatase from rat liver mitochondria. 3. Subunit composition. J Biol Chem 248, 7427-7431, PubMed ID: 74012057 Feo, F. et al. (1975) Effect of cholesterol content on some physical and functional properties of mitochondria isolated from adult rat liver, fetal liver, cholesterol-enriched liver and hepatomas AH-130, 3924A and 5123. Biochim Biophys Acta 413, 116-134, PubMed ID: 76062570 Parlo, R.A. and Coleman, P.S. (1984) Enhanced rate of citrate export from cholesterol-rich

in molecular medicine

33

34

35

36

37

38

39

40

41

42

43

44

45

(1975) On the DNA polymerase III of mouse myeloma: partial purification and characterization. Biochemistry 14, 1006-1020, PubMed ID: 75146411 Shay, J.W. and Werbin, H. (1987) Are mitochondrial DNA mutations involved in the carcinogenic process? Mutat Res 186, 149-160, PubMed ID: 87315142 Torri, A.F., Kunkel, T.A. and Englund, P.T. (1994) A beta-like DNA polymerase from the mitochondrion of the trypanosomatid Crithidia fasciculata. J Biol Chem 269, 8165-8171, PubMed ID: 94179191 Grossman, L.I. and Shoubridge, E.A. (1996) Mitochondrial genetics and human disease. Bioessays 18, 983-991, PubMed ID: 97130392 Johns, D.R. (1995) Seminars in medicine of the Beth Israel Hospital, Boston. Mitochondrial DNA and disease. N Engl J Med 333, 638-644, PubMed ID: 95364863 Naviaux, R.K. (2000) Mitochondrial DNA disorders. Eur J Pediatr 159 Suppl 3, S219-226, PubMed ID: 21084744 White, A.J. (2001) Mitochondrial toxicity and HIV therapy. Sex Transm Infect 77, 158-173, PubMed ID: 21295616 Pedersen, P.S. et al. (1999) Ion transport in epithelial spheroids derived from human airway cells. Am J Physiol 276, L75-80, PubMed ID: 99103968 Dang, C.V. et al. (1997) Oncogenes in tumor metabolism, tumorigenesis, and apoptosis. J Bioenerg Biomembr 29, 345-354, PubMed ID: 98048336 Lundholm, K. et al. (1982) Glucose turnover, gluconeogenesis from glycerol, and estimation of net glucose cycling in cancer patients. Cancer 50, 1142-1150, PubMed ID: 82258815 Mazurek, S., Boschek, C.B. and Eigenbrodt, E. (1997) The role of phosphometabolites in cell proliferation, energy metabolism, and tumor therapy. J Bioenerg Biomembr 29, 315-330, PubMed ID: 98048333 Fischer, C.P., Bode, B.P. and Souba, W.W. (1998) Adaptive alterations in cellular metabolism with malignant transformation. Ann Surg 227, 627634; discussion 634-626, PubMed ID: 98266691 Ockner, R.K., Kaikaus, R.M. and Bass, N.M. (1993) Fatty-acid metabolism and the pathogenesis of hepatocellular carcinoma: review and hypothesis. Hepatology 18, 669-676, PubMed ID: 93366320 Shaw, J.H. and Wolfe, R.R. (1987) Fatty acid and

46 47

48

49

50 51 52

53

54

55

56

57

58

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

14

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews
mitochondrial NADH-ubiquinone reductase activity via a rotenone-type mechanism by two of the dyes. Biochem Pharmacol 45, 691-696, PubMed ID: 93183171 Bleday, R. et al. (1986) Inhibition of rat colon tumor isograft growth with dequalinium chloride. Arch Surg 121, 1272-1275, PubMed ID: 87048126 Weiss, M.J. et al. (1987) Dequalinium, a topical antimicrobial agent, displays anticarcinoma activity based on selective mitochondrial accumulation. Proc Natl Acad Sci U S A 84, 54445448, PubMed ID: 87261002 Koya, K. et al. (1996) MKT-077, a novel rhodacyanine dye in clinical trials, exhibits anticarcinoma activity in preclinical studies based on selective mitochondrial accumulation. Cancer Res 56, 538-543, PubMed ID: 96147266 Sun, X. et al. (1994) AA1, a newly synthesized monovalent lipophilic cation, expresses potent in vivo antitumor activity. Cancer Res 54, 1465-1471, PubMed ID: 94184962 Rideout, D., Bustamante, A. and Patel, J. (1994) Mechanism of inhibition of FaDu hypopharyngeal carcinoma cell growth by tetraphenylphosphonium chloride. Int J Cancer 57, 247-253, PubMed ID: 94208957 Modica-Napolitano, J.S. et al. (1996) Selective damage to carcinoma mitochondria by the rhodacyanine MKT-077. Cancer Res 56, 544-550, PubMed ID: 96147267 Weisberg, E.L. et al. (1996) In vivo administration of MKT-077 causes partial yet reversible impairment of mitochondrial function. Cancer Res 56, 551-555, PubMed ID: 96147268 Lampidis, T.J. et al. (1984) Effects of the mitochondrial probe rhodamine 123 and related analogs on the function and viability of pulsating myocardial cells in culture. Agents Actions 14, 751-757, PubMed ID: 84303640 Modica-Napolitano, J.S. et al. (1990) Mitochondrial toxicity of cationic photosensitizers for photochemotherapy. Cancer Res 50, 7876-7881, PubMed ID: 91070553 Dougherty, T.J., Weishaupt, K.R. and Boyle, D.G. (1985) Photodynamic Sensitizers, J.B. Lipincott Co., Philadelphia, PA, USA Wilson, B.C. and Jeeves, W.P. (1987) Photodynamic therapy of cancer. In Photomedicine (Vol. 2) (Ben-Hur, E. and Rosenthal, I., eds), pp. 127-177, CRC Press, Boca Raton, FL, USA Powers, S.K. (1988) Photochemotherapy. In

in molecular medicine

59

60

61

62

63

64

65

66

67

68

69

70

hepatoma mitochondria. The truncated Krebs cycle and other metabolic ramifications of mitochondrial membrane cholesterol. J Biol Chem 259, 9997-10003, PubMed ID: 84289473 Costello, L.C. and Franklin, R.B. (2000) The intermediary metabolism of the prostate: a key to understanding the pathogenesis and progression of prostate malignancy. Oncology 59, 269-282, PubMed ID: 20549295 Summerhayes, I.C. et al. (1982) Unusual retention of rhodamine 123 by mitochondria in muscle and carcinoma cells. Proc Natl Acad Sci U S A 79, 5292-5296, PubMed ID: 83039355 Johnson, L.V. et al. (1981) Monitoring of relative mitochondrial membrane potential in living cells by fluorescence microscopy. J Cell Biol 88, 526535, PubMed ID: 81168401 Bernal, S.D. et al. (1982) Rhodamine-123 selectively reduces clonal growth of carcinoma cells in vitro. Science 218, 1117-1119, PubMed ID: 83067410 Lampidis, T.J. et al. (1983) Selective toxicity of rhodamine 123 in carcinoma cells in vitro. Cancer Res 43, 716-720, PubMed ID: 83076897 Bernal, S.D. et al. (1983) Anticarcinoma activity in vivo of rhodamine 123, a mitochondrial- specific dye. Science 222, 169-172, PubMed ID: 84017536 Modica-Napolitano, J.S. and Aprille, J.R. (1987) Basis for the selective cytotoxicity of rhodamine 123. Cancer Res 47, 4361-4365, PubMed ID: 87273236 Nicholls, D.G. (1981) Bioenergetics: an Introduction to the Chemiosmotic Theory, Academic Press, New York, USA Davis, S. et al. (1985) Mitochondrial and plasma membrane potentials cause unusual accumulation and retention of rhodamine 123 by human breast adenocarcinoma-derived MCF-7 cells. J Biol Chem 260, 13844-13850, PubMed ID: 86033856 Modica-Napolitano, J.S. et al. (1984) Rhodamine 123 inhibits bioenergetic function in isolated rat liver mitochondria. Biochem Biophys Res Commun 118, 717-723, PubMed ID: 84153846 Emaus, R.K., Grunwald, R. and Lemasters, J.J. (1986) Rhodamine 123 as a probe of transmembrane potential in isolated rat- liver mitochondria: spectral and metabolic properties. Biochim Biophys Acta 850, 436-448, PubMed ID: 86269938 Anderson, W.M. et al. (1993) Cytotoxic effect of thiacarbocyanine dyes on human colon carcinoma cells and inhibition of bovine heart

71

72

73

74

75

76

77

78

79

80

81

82

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

15

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews
94 Capuano, F. et al. (1996) Oxidative phosphorylation and F(O)F(1) ATP synthase activity of human hepatocellular carcinoma. Biochem Mol Biol Int 38, 1013-1022, PubMed ID: 97013420 95 Cuezva, J.M. et al. (1997) Mitochondrial biogenesis in the liver during development and oncogenesis. J Bioenerg Biomembr 29, 365-377, PubMed ID: 98048338 96 Modica-Napolitano, J.S. and Touma, S.E. (2000) Functional differences in mitochondrial enzymes from normal epithelial and carcinoma cells. In Mitochondrial Dysfunction in Pathogenesis, A Keystone Symposium (1520 January, Santa Fe, USA), p. 64, Keystone Symposia, Silverthorne, CO 80498, USA 97 Sun, A.S., Sepkowitz, K. and Geller, S.A. (1981) A study of some mitochondrial and peroxisomal enzymes in human colonic adenocarcinoma. Lab Invest 44, 13-17, PubMed ID: 81097387 98 Sun, A.S. and Cederbaum, A.I. (1980) Oxidoreductase activities in normal rat liver, tumor-bearing rat liver, and hepatoma HC252. Cancer Res 40, 4677-4681, PubMed ID: 81064485 99 Heerdt, B.G. et al. (1990) Expression of mitochondrial cytochrome c oxidase in human colonic cell differentiation, transformation, and risk for colonic cancer. Cancer Res 50, 1596-1600, PubMed ID: 90150125 100 Luciakova, K. and Kuzela, S. (1992) Increased steady-state levels of several mitochondrial and nuclear gene transcripts in rat hepatoma with a low content of mitochondria. Eur J Biochem 205, 1187-1193, PubMed ID: 92249325 101 Costantini, P. et al. (2000) Mitochondrion as a novel target of anticancer chemotherapy. J Natl Cancer Inst 92, 1042-1053, PubMed ID: 20341836 102 Chan, S.H. and Barbour, R.L. (1983) Adenine nucleotide transport in hepatoma mitochondria. Characterization of factors influencing the kinetics of ADP and ATP uptake. Biochim Biophys Acta 723, 104-113, PubMed ID: 83153600 103 Sul, H.S. et al. (1979) Comparison of the adenine nucleotide translocase in hepatomas and rat liver mitochondria. Biochim Biophys Acta 551, 148155, PubMed ID: 79145476 104 Woldegiorgis, G. and Shrago, E. (1985) Adenine nucleotide translocase activity and sensitivity to inhibitors in hepatomas. Comparison of the ADP/ATP carrier in mitochondria and in a purified reconstituted liposome system. J Biol Chem 260, 7585-7590, PubMed ID: 85207805

in molecular medicine

83

84

85

86

87

88

89

90

91

92

93

Application of Lasers in Neurosurgery (Cerullo, L.J., ed.), pp. 137-155, Year Book Medical Publishers, Inc., Chicago, IL, USA Kowaltowski, A.J. et al. (1999) Mitochondrial effects of triarylmethane dyes. J Bioenerg Biomembr 31, 581-590, PubMed ID: 20145121 Oseroff, A.R. et al. (1987) Strategies for selective cancer photochemotherapy: antibody-targeted and selective carcinoma cell photolysis. Photochem Photobiol 46, 83-96, PubMed ID: 87290165 Powers, S.K. et al. (1986) Laser photochemotherapy of rhodamine-123 sensitized human glioma cells in vitro. J Neurosurg 64, 918923, PubMed ID: 86198943 Beckman, W.C., Jr. et al. (1987) Differential retention of rhodamine 123 by avian sarcoma virus-induced glioma and normal brain tissue of the rat in vivo. Cancer 59, 266-270, PubMed ID: 87102460 Powers, S.K. et al. (1989) Photosensitization of human glioma cells by chalcogenapyrylium dyes. J Neurooncol 7, 179-188, PubMed ID: 89381799 Oseroff, A.R. et al. (1986) Intramitochondrial dyes allow selective in vitro photolysis of carcinoma cells. Proc Natl Acad Sci U S A 83, 9729-9733, PubMed ID: 87092321 Ara, G. et al. (1987) Mechanisms of mitochondrial photosensitization by the cationic dye, N,N- bis(2-ethyl-1,3dioxylene)kryptocyanine (EDKC): preferential inactivation of complex I in the electron transport chain. Cancer Res 47, 6580-6585, PubMed ID: 88052644 Walstad, D.L., Brown, J.T. and Powers, S.K. (1989) The effect of a chalcogenapyrylium dye with and without photolysis on mitochondrial function in normal and tumor cells. Photochem Photobiol 49, 285-291, PubMed ID: 89283000 Modica-Napolitano, J.S. et al. (1998) Photoactivation enhances the mitochondrial toxicity of the cationic rhodacyanine MKT-077. Cancer Res 58, 71-75, PubMed ID: 98086029 Pedersen, P.L. and Morris, H.P. (1974) Uncouplerstimulated adenosine triphosphatase activity. Deficiency in intact mitochondria from Morris hepatomas and ascites tumor cells. J Biol Chem 249, 3327-3334, PubMed ID: 74172590 Capuano, F., Guerrieri, F. and Papa, S. (1997) Oxidative phosphorylation enzymes in normal and neoplastic cell growth. J Bioenerg Biomembr 29, 379-384, PubMed ID: 98048339

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

16

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews
the microsatellite mutator phenotype. Science 275, 967-969, PubMed ID: 97172557 Brimmell, M. et al. (1998) BAX frameshift mutations in cell lines derived from human haemopoietic malignancies are associated with resistance to apoptosis and microsatellite instability. Oncogene 16, 1803-1812, PubMed ID: 98243036 Singh, K.K. et al. (1999) Mitochondrial DNA determines the cellular response to cancer therapeutic agents. Oncogene 18, 6641-6646, PubMed ID: 20065137 Weissig, V. and Torchilin, V.P., eds (2001) Drug and DNA delivery to mitochondria. Adv Drug Deliv Rev 49, 1-2, PubMed ID: 21271658 Rizzuto, R. et al. (1992) Rapid changes of mitochondrial Ca2+ revealed by specifically targeted recombinant aequorin. Nature 358, 325327, PubMed ID: 92350252 Ellerby, H.M. et al. (1999) Anti-cancer activity of targeted pro-apoptotic peptides. Nat Med 5, 1032-1038, PubMed ID: 99401083 Clayton, D.A. and Vinograd, J. (1967) Circular dimer and catenate forms of mitochondrial DNA in human leukaemic leucocytes. J Pers 35, 652657, PubMed ID: 68135375 Clayton, D.A. and Vinograd, J. (1969) Complex mitochondrial DNA in leukemic and normal human myeloid cells. Proc Natl Acad Sci U S A 62, 1077-1084, PubMed ID: 69244656 Gamen, S. et al. (1995) mtDNA-depleted U937 cells are sensitive to TNF and Fas-mediated cytotoxicity. FEBS Lett 376, 15-18, PubMed ID: 96096778 Boultwood, J. et al. (1996) Amplification of mitochondrial DNA in acute myeloid leukaemia. Br J Haematol 95, 426-431, PubMed ID: 97060858 Ivanova, R. et al. (1998) Mitochondrial DNA sequence variation in human leukemic cells. Int J Cancer 76, 495-498, PubMed ID: 98250006 Hudson, B. and Vinograd, J. (1967) Catenated circular DNA molecules in HeLa cell mitochondria. Nature 216, 647-652, PubMed ID: 68135373 Welter, C. et al. (1989) Alteration of mitochondrial DNA in human oncocytomas. Genes Chromosomes Cancer 1, 79-82, PubMed ID: 91120367 Tallini, G. (1998) Oncocytic tumours. Virchows Arch 433, 5-12, PubMed ID: 98355568 Tallini, G. et al. (1994) Analysis of nuclear and mitochondrial DNA alterations in thyroid and renal oncocytic tumors. Cytogenet Cell Genet 66,

in molecular medicine

105 Halestrap, A.P. and Davidson, A.M. (1990) Inhibition of Ca2(+)-induced large-amplitude swelling of liver and heart mitochondria by cyclosporin is probably caused by the inhibitor binding to mitochondrial-matrix peptidyl-prolyl cis-trans isomerase and preventing it interacting with the adenine nucleotide translocase. Biochem J 268, 153-160, PubMed ID: 90262539 106 Bernardi, P. et al. (1998) The mitochondrial permeability transition. Biofactors 8, 273-281, PubMed ID: 99113258 107 Faure Vigny, H. et al. (1996) Expression of oxidative phosphorylation genes in renal tumors and tumoral cell lines. Mol Carcinog 16, 165-172, PubMed ID: 96313221 108 Giraud, S. et al. (1998) Expression of human ANT2 gene in highly proliferative cells: GRBOX, a new transcriptional element, is involved in the regulation of glycolytic ATP import into mitochondria. J Mol Biol 281, 409-418, PubMed ID: 98365507 109 Reed, J.C. (1994) Bcl-2 and the regulation of programmed cell death. J Cell Biol 124, 1-6, PubMed ID: 94124601 110 Kroemer, G. (1997) The proto-oncogene Bcl-2 and its role in regulating apoptosis. Nat Med 3, 614620, PubMed ID: 97319583 111 Reed, J.C. (1997) Double identity for proteins of the Bcl-2 family. Nature 387, 773-776, PubMed ID: 97337864 112 Venturini, I. et al. (1998) Up-regulation of peripheral benzodiazepine receptor system in hepatocellular carcinoma. Life Sci 63, 1269-1280, PubMed ID: 98442858 113 Galiegue, S. et al. (1999) Cloning and characterization of PRAX-1. A new protein that specifically interacts with the peripheral benzodiazepine receptor. J Biol Chem 274, 29382952, PubMed ID: 99115641 114 OGorman, E. et al. (1997) The role of creatine kinase in inhibition of mitochondrial permeability transition. FEBS Lett 414, 253-257, PubMed ID: 97459736 115 Schiemann, S. et al. (1998) Molecular analysis of two mammary carcinoma cell lines at the transcriptional level as a model system for progression of breast cancer. Clin Exp Metastasis 16, 129-139, PubMed ID: 98173027 116 Beurdeley-Thomas, A. et al. (2000) The peripheral benzodiazepine receptors: a review. J Neurooncol 46, 45-56, PubMed ID: 20353100 117 Rampino, N. et al. (1997) Somatic frameshift mutations in the BAX gene in colon cancers of

118

119

120

121

122

123

124

125

126

127

128

129

130 131

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

17

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews
various types of liver disease and in normal liver. Hepatology 21, 1547-1551, PubMed ID: 95286138 El Meziane, A. et al. (1998) Mitochondrial tRNALeu isoforms in lung carcinoma cybrid cells containing the np 3243 mtDNA mutation. Hum Mol Genet 7, 2141-2147, PubMed ID: 99036686 Savre-Train, I, Piatyszek, M.A. and Shay, J.W. (1992) Transcription of deleted mitochondrial DNA in human colon adenocarcinoma cells. Hum Mol Genet 1, 203-204, PubMed ID: 93265014 Burgart, L.J. et al. (1995) Somatic mitochondrial mutation in gastric cancer. Am J Pathol 147, 11051111, PubMed ID: 96010274 Liang, B.C. (1996) Evidence for association of mitochondrial DNA sequence amplification and nuclear localization in human low-grade gliomas. Mutat Res 354, 27-33, PubMed ID: 96305473 Richard, S.M. et al. (2000) Nuclear and mitochondrial genome instability in human breast cancer. Cancer Res 60, 4231-4237, PubMed ID: 20399732

in molecular medicine

253-259, PubMed ID: 94215316 132 Heddi, A. et al. (1996) Coordinate expression of nuclear and mitochondrial genes involved in energy production in carcinoma and oncocytoma. Biochim Biophys Acta 1316, 203209, PubMed ID: 96375169 133 Kovacs, A. et al. (1992) Mitochondrial and chromosomal DNA alterations in human chromophobe renal cell carcinomas. J Pathol 167, 273-277, PubMed ID: 92389115 134 Horton, T.M. et al. (1996) Novel mitochondrial DNA deletion found in a renal cell carcinoma. Genes Chromosomes Cancer 15, 95-101, PubMed ID: 96431090 135 Selvanayagam, P. and Rajaraman, S. (1996) Detection of mitochondrial genome depletion by a novel cDNA in renal cell carcinoma. Lab Invest 74, 592-599, PubMed ID: 96175616 136 Yamamoto, H. et al. (1992) Significant existence of deleted mitochondrial DNA in cirrhotic liver surrounding hepatic tumor. Biochem Biophys Res Commun 182, 913-920, PubMed ID: 92134320 137 Fukushima, S. et al. (1995) The frequency of 4977 base pair deletion of mitochondrial DNA in

138

139

140

141

142

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

18

Mitochondria as targets for detection and treatment of cancer

http://www-ermm.cbcu.cam.ac.uk

expert reviews

in molecular medicine

Further reading, resources and contacts


Wallace, D.C. (1997) Mitochondrial DNA in aging and disease. Sci Am 277, 40-47, PubMed ID: 97388606 Singh, K.K. (1998) Mitochondrial DNA Mutation in Aging, Disease and Cancer, Springer, New York Scheffler, I.E. (1999) Mitochondria, Wiley-Liss, New York Naviaux, R.K. (1997) Spectrum of Mitochondrial Diseases, Exceptional Parent, http:// biochemgen.ucsd.edu/mmdc/ep-toc.htm The Mitochondria Research Society website provides basic information and useful links on mitochonrial disease for scientists, clinicians and patients. http://www.mitoresearch.org Website providing introduction to mitochondrial DNA for students (The fire within: the unfolding story of human mtDNA): http://biocrs.biomed.brown.edu/books/essays/mitochondrialDNA.html The Mitomap database lists polymorphisms and mutations of human mitochondrial DNA. http://www.gen.emory.edu/mitomap.html Keshav K. Singh home page: http://www.jhmi.edu/~kksingh

Features associated with this article Figures


Figure 1. Schematic of key mitochondrial metabolic pathways (fig001ksb). Figure 2. Mitochondria of neuron revealed by staining with a rhodamine 123 derivative (fig002ksb).

Tables
Table 1. Mitochondrial diseases (tab001ksb). Table 2. Mitochondrial DNA mutation in cancers (tab002ksb).

Citation details for this article


Josephine S. Modica-Napolitano and Keshav K. Singh (2002) Mitochondria as targets for detection and treatment of cancer. Exp. Rev. Mol. Med. 11 April, http://www-ermm.cbcu.cam.ac.uk/02004453h.htm

Accession information: (02)00445-3a.pdf (short code: txt001ksb); 11 April 2002 ISSN 1462-3994 2002 Cambridge University Press

19

Mitochondria as targets for detection and treatment of cancer

You might also like