Zum Hauptinhalt springen

Rescue of TCA Cycle Dysfunction for Cancer Therapy

Marquez, Jubert ; Ah, Kim ; et al.
In: Journal of Clinical Medicine, Jg. 8 (2019-12-06), S. 2161-2161
Online unknown

Rescue of TCA Cycle Dysfunction for Cancer Therapy 

Mitochondrion, a maternally hereditary, subcellular organelle, is the site of the tricarboxylic acid (TCA) cycle, electron transport chain (ETC), and oxidative phosphorylation (OXPHOS)—the basic processes of ATP production. Mitochondrial function plays a pivotal role in the development and pathology of different cancers. Disruption in its activity, like mutations in its TCA cycle enzymes, leads to physiological imbalances and metabolic shifts of the cell, which contributes to the progression of cancer. In this review, we explored the different significant mutations in the mitochondrial enzymes participating in the TCA cycle and the diseases, especially cancer types, that these malfunctions are closely associated with. In addition, this paper also discussed the different therapeutic approaches which are currently being developed to address these diseases caused by mitochondrial enzyme malfunction.

Keywords: mitochondria; TCA; cancer; IDH; SDH; FH; MDH; CRISPR/Cas9; miRNA

1. Introduction

Mitochondrion is a maternally hereditary, subcellular organelle which plays a role in bioenergetics, biosynthesis, and cell signaling [[1]]. The human mitochondrial proteome is composed of a subset of ~20,000 distinct mammalian proteins which are localized in the said organelle. Thirteen of these proteins are encoded by mitochondrial DNA (mtDNA) and the rest are encoded by nuclear DNA (nDNA) [[2]]. Mutations in these mitochondrial protein genes are notably implicated in diseases such as cancer and diabetes, as well as a plethora of other genetic diseases [[3]]. Mitochondria have a double lipid membrane with various types of membrane proteins which are divided into four divisions: The intermembrane space (IMS), outer mitochondrial membrane (OMM), inner mitochondrial membrane (IMM), which has a highly particular structure to create cristae of large surface area for ATP production, and the mitochondrial matrix [[4]]. The mitochondria play a key role in ATP production and circulation based on the availability of energy from calories and oxygen, along with the demands for cellular maintenance and reproduction [[3]]. Carbon sources from glycolysis, fatty acids, and glutamine are utilized to produce ATP. Carbon sources entering the tricarboxylic acid (TCA) cycle in the mitochondrial matrix produce NADH and FADH2, which transfer their electrons to the electron transport chain (ETC) located in the IMM [[5]]. In the ETC system, the electrons transferred from NADH/FADH2 to oxygen induce an oxidation-reduction reaction at each step, and energy from the oxidized electron is utilized to pump protons from the mitochondrial matrix into the intermembrane space through complex I (NADH dehydrogenase), complex III (CoQH2-cytochrome c reductase), and complex IV (cytochrome c oxidase) [[6]]. The proton gradient is harnessed to drive the switch of ADP to ATP by complex V (ATP synthase), during which concurrently pumped protons return to the matrix. In the matrix, lipids are oxidized by β-oxidation as a breakdown of fatty chains to produce acetyl-CoA [[7]].

Energy production during the metabolic process has recently been in the spotlight due to its capability to generate signaling molecules for various cellular responses. Recent technological developments have further supported long-withstanding hypotheses regarding the key role of aberrant energy production and metabolism in disease models (Figure 1). Therefore, in this review, we discussed the metabolic differences in normal and disease models, and highlighted TCA enzymes and proteins critical in further understanding disease progression along with how we can harness the knowledge regarding these enzymes and proteins in order to address diseases, especially in cancer.

2. The TCA Cycle: In Sickness and in Health

The TCA cycle unifies the carbohydrate, lipid, and protein metabolism pathways. In a healthy, normal cell, glycolysis is responsible for the oxidation of the glucose molecule into pyruvate to produce ATP, which is then decarboxylated into acetyl-CoA as it enters the mitochondria, allowing it to enter the TCA cycle. However, each body part exhibits strikingly different metabolic profiles. For example, in the human brain, with the exception of prolonged fasting states, glucose is the main source of energy [[8]]. Muscles, on the other hand, have a vast reservoir of glycogen that can easily be converted into glucose 6-phosphate [[9]]. The liver, which is controlled by both neuronal and hormonal systems, provides energy for organs, such as the brain and muscle, in addition to extrahepatic tissues. The liver can produce glucose by breaking down its stored glycogen and through gluconeogenesis. However, in the fasting state, ketone bodies' conversion from fatty acids is facilitated through mitochondrial β-oxidation and ketogenesis [[10]].

TCA intermediates also play important roles in pathways in which they leave the cycle to be converted into glucose, fatty acids, or non-essential amino acids. Once removed, these intermediates need to be replaced to allow continued function and cycle, known as anaplerosis [[12]]. In heart and skeletal tissues, anaplerosis maintains steady-state concentrations of TCA intermediates [[13]]. Cancer cells, on the other hand, transport glucose-derived pyruvate into the mitochondria, where it is used as an anaplerotic substrate to replace TCA intermediates used for biosynthesis [[14]]. Inadequate amounts of glutamine or suppressed glutaminase forces the cancer cells to depend on glucose carbon flux through pyruvate carboxylase to keep oxaloacetate production and continue downstream TCA cycle activity [[15]]. In cases like non-small-cell lung carcinoma and glioblastoma, they more frequently rely on pyruvate anaplerosis to maintain TCA cycle flux.

In comparison to the organs previously mentioned, there is an immense demand for energy in mammalian hearts due to its continuous and incessant beating. Central to energy transduction in the heart, the mitochondria generate more than 95% of the ATP used by the heart. In a normal heart, fatty acyl-coenzyme A (CoA) and pyruvate fuel the mitochondria. The entry of long-chain acyl-CoA into the mitochondria is rate-limited by carnitine-palmitoyl transferase-1 (CPT1), while pyruvate dehydrogenase (PDH) reaction regulates pyruvate oxidation. Substrates, such as lactate, ketone bodies, and amino acids, freely enter the mitochondria for oxidation [[16]].

3. TCA Enzymes: The Future to Understanding The Complexities of Diseases

Diseases of the TCA cycle constitute a group of rare human diseases that affect core mitochondrial metabolism [[17]]. The deficiency of enzymes involving the TCA cycle was detected in order to obtain crucial roles in several human diseases.

Emerging evidences suggest that cancer is mitochondrial in nature [[18]]. Warburg originally observed and postulated that excess lactate production by tumors in the presence of oxygen is a sign of mitochondrial dysfunction, eventually giving rise to the idea of aerobic glycolysis or the 'Warburg effect' [[20]]. Recent studies proved that the dysfunction observed by Warburg is merely an altered state of the mitochondria and is a part of a bigger picture in cancer bioenergetics [[21]]. The majority of cancer cells generate most of their ATP through the mitochondria. Few tumors bear TCA enzymes mutations, such as isocitrate dehydrogenase (IDH), succinate dehydrogenase (SDH), fumarate hydratase (FH), and malate dehydrogenase (MDH) [[24]]. Their mutations are also involved directly or indirectly, which comprises the activation of a hypoxic cellular response and high levels of ROS often found in cancer cells [[25]].

3.1. Isocitrate Dehydrogenase

Isocitrate dehydrogenase (IDH) is mainly known for its role in catalyzing the oxidative decarboxylation of isocitrate, resulting in 2-oxoglutarate (α-KG) and CO2 (Figure 2). IDH exists in three isoforms: IDH1 is present in the cytoplasm and peroxisomes, while IDH2 and IDH3 are located in the mitochondrial matrix. IDH1/2 isoforms were also identified to mediate the reverse reductive carboxylation of α-KG to isocitrate, which oxidizes NADPH to NADP+. Meanwhile, IDH3 only facilitates the irreversible, NAD-dependent conversion of isocitrate to α-KG [[26]]. ODH3 mutation of IDH1/2 is found in low-grade glioma and secondary glioblastoma (GBM), chondrosarcoma, intrahepatic cholangiocarcinomas, hematologic malignancies, premalignant diseases, and rare inherited metabolism disorders [[28]]. In a clinical study, IDH1 and IDH2 mutations were observed in 16%–17% of patients with AML, in around 20% of angioimmunoblastic T-cell lymphomas (AITL) with worse prognosis [[29]], and in some low-frequency cancer malignancies [[30]]. Mutations in the gene-encoding the said enzymes cause increased production of R-2-hydroxyglutarate (R-2HG), an oncogenic factor promoting leukemogenesis (Figure 3) [[34]]. R-2HG produced by mutant IDH in low-grade glioma was shown to activate the mammalian target of rapamycin (mTOR) signaling pathway, which is important for cell growth and metabolism [[36]]. Moreover, inhibition of the IDH2/R140Q somatic mutant inhibitor induces differentiation of the human erythroleukemic cell line (HEL) and human primary AML cells [[37]]. Furthermore, mutant IDH induces hypermethylation of MIRNA148A, a tumor-suppressive miRNA in glioma CpG island methylator phenotype (G-CIMP) [[38]].

The most common IDH mutation found in cancer is the substitution of a single arginine in the catalytic site of the enzyme, R132 in IDH1 and R140 or R172 in IDH2, which results in a gain of function. Alterations in R132 in IDH1 and either R172 or R140 in IDH2 represent the majority of IDH mutations identified in cancers [[39]]. IDH1/R132 and IDH2/R172 are commonly found in gliomas, cholangiocarcinomas, and chondrosarcomas, with a higher frequency of IDH1/R132 mutation occurring in these cancers (58%–90%, 40%–50%, and 50%–60% respectively) compared to IDH2/R172 (3%–5%, 5%–10%, and 10%). These mutations are also found at relatively lower frequencies in AML. Meanwhile, IDH2/R140 is the most common mutation found in AML (30%–50%). However, IDH2/R140, unlike IDH1/R132 and IDH2/R172 mutations, is not found in gliomas, cholangiocarcinomas, and chondrosarcomas [[40]]. IDH2/R140Q is the most common mutation (75%–80%) and confers a favorable or insignificant impact on overall survival [[41]]. However, IDH2/R172K mutation is found in 20% of the cases, with a lower complete remission rate, higher relapse rate, and lower overall survival [[46]]. Three variants that occur in exon 4 were discovered upon IDH2 gene screening: c.543+45G>A, c. 389 A>T, p. Lys120Met and c.414 T>C, and p.Thr138Thr. These gene variants were found in two independent patients classified under French and Tunisian familial cases which, despite ethnic differences, were similarly diagnosed with non-Hodgkin lymphoma [[48]]. Melissa Carbonneau et al. found that the molecular mechanism underlying the oncogenic activity of mutant IDH1/2 involved mTOR signaling via KDM4A inhibition, an αKG-dependent enzyme [[35]].

3.2. Succinate Dehydrogenase

Succinate dehydrogenase (SDH) is an enzyme bound to the inner mitochondrial membrane, where it oxidizes succinate to fumarate, and is classified as a tumor suppressor [[51]]. The SDH complex consists of four subunits (SDHA, SDHB, SDHC, and SDHD), and a deficiency of this enzyme is known to activate tumor formation through dysregulation of HIF activity [[52]]. HIF, which is a transcription factor, can activate anti-apoptotic and pro-proliferation genes, leading to tumor formation of cancer cells (Figure 3). In addition, deficiency of the enzyme can cause accumulation of the metabolite, succinate. Succinate was shown to exert pro-inflammatory effects through the generation of mtROS [[53]]. The SDH complex gene-associated cancers include paragangliomas, pheochromocytomas, gastrointestinal stromal tumors (GIST), SDH-deficient renal cell carcinoma [[54]], thyroid tumors, neuroblastomas, testicular seminoma, and ovarian cancer [[56]].

The SDHA gene is responsible for encoding SDH enzyme major catalytic subunit, possessing a covalently attached flavin adenine dinucleotide (FAD) prosthetic group which can bind with substrates, such as fumarate and succinate, and also with physiological regulators, such as oxaloacetate and ATP. Inactivation of SDHA has been shown to promote neurodegenerative diseases like Leigh syndrome, which is an early-onset encephalopathy [[57]], as well as late-onset optic atrophy, ataxia, and myopathy [[60]]. In addition, a missense mutation in SDHA was shown to cause a multisystemic failure, leading to neonatal death [[61]]. Meanwhile, in SDH-deficient GISTs, despite the presence of other SDH subunit gene germline mutations, SDHA is the most prevalent among the four [[62]].

SDHB is an enzyme that catalyzes succinate oxidation. SDHB mutations normally lead to extra-adrenal paragangliomas (PGLs), which are usually characterized by highly aggressive tumors, poor prognosis, and early-age onset (~30 years) [[63]]. To a lesser extent, it may also impose risks of adrenal pheochromocytoma (PCC) and head and neck paragangliomas (HNPGLs) [[63]]. In addition, renal cell carcinoma and T-cell acute leukemia are also associated with SDHB mutations [[67]]. In the study of Fishbein et al., the authors collected and screened data of 173 PGLs/PCCs patients from The Cancer Genome Atlas. SDHB appeared to be the most common germline mutation (9%) and exhibited the highest number of copy number alteration (57%) [[69]].

SDHC mutations were initially implicated with HNPGLs alone. However, recent rare cases of adrenal PCCs and extra-adrenal PGLs were observed to be related to SDHC mutation as well [[70]]. Clinically, features of SDHC-associated cases are similar to symptoms of sporadic HNPGLs [[73]]. In addition, somatic SDHC mutations were also detected in 5% of sporadic thyroid cancer cases in a cohort study [[74]].

SDHD mutations are usually related to multifocal HNPGLs and sometimes to adrenal PCCs and extra-adrenal PGLs, which are usually benign. Metastatic HNPGLs have been described within SDHD mutation carriers with 0%–10% prevalence [[70], [75]]. In the study of Marc Bennedbaek et al., the authors identified 18 different germline variants of SDH in the Danish PGL and PCC patients, wherein 12 were likely pathogenic/pathogenic [[76]]. Furthermore, PGL/PCC syndrome has also been associated with mutations in SDH assembly factor 2 (SDHAF2) [[77]], which is required for the flavination of SDH [[79]]. Meanwhile, in sporadic thyroid cancers, about 6% of patients showed germline mutation of SDHB or SDHD [[74]].

3.3. Fumarate Hydratase

Fumarate hydratase (FH) is responsible for the hydration/dehydration of fumarate to malate, an integral process in cellular respiration and energy production. Similarly to succinate, an increase in fumarate inhibits prolyl hydroxylases, which are responsible for the regulation of HIF-1α degradation. Fumarate upregulation may also cause post-translational modification and inactivation of Kelch-like ECH-associated protein 1 (KEAP1). KEAP1 is a substrate adapter protein for the E3 ubiquitin ligase complex which targets nuclear factor erythroid 2-related factor (NRF2) [[80]]. NRF2, on the other hand, is a regulator of cellular antioxidant defense [[81]]. Fumarate have also been proved to bind with glutathione to form the oncometabolite succinate glutathione (GSF), which can act as an alternative substrate to glutathione reductase, thus decreasing NADPH levels and enhancing mt ROS and HIF-1 activation. The aforementioned binding can also cause a further increase in oxidative stress due to the depletion of the antioxidant molecules in the system. Furthermore, in the study by Tyrakis et al., increased fumarate due to the deficiency of FH caused the impairment of the respiratory chain complex 2 via the succination of members of Fe-S cluster biogenesis proteins, which are important for the activity of mitochondrial enzymes [[82]].

Fumarase gene germline mutation is connected to multiple cutaneous and uterine leiomyomas (MCUL) and hereditary leiomyomatosis and renal cell cancer (HLRCC) [[83]]. Meanwhile, in a study where tissue samples from leiomyosarcoma and uterine leiomyoma patients were analyzed, no somatic mutations in the fumarate gene was detected, implying that the somatic mutation in gene-encoding the said enzyme does not play a major role in the development of sporadic leiomyosarcomas or uterine leiomyomas [[85]]. On the other hand, in the cohort study of PGLs conducted by Letouzé et al., FH somatic mutation was detected in the only sample of hypermethylated PGL that did not possess SDHx mutation [[86]].

An estimated 90% (76%–100%) of families were found to have clinically suggestive HLRCC with predisposed early onset, aggressive form of type 2 papillary renal cell carcinoma [[87]]. FH mutation in kidney cancer has been shown to induce an increase in glucose uptake, glycolytic rate, and contribution of glucose to the pentose phosphate pathway [[89]]. In another study in clear cell renal cancer, a mutation in FH led to the accumulation of HIF-2α, a promotor renal carcinogenesis [[90]]. However, Tong. W.H. et al. showed that FH mutations in kidney cancer are associated with a reduction in the activity of the metabolic sensor, AMP-activated protein kinase (AMPK), which leads to increased synthesis of fatty acids and proteins to support ongoing cellular anabolism [[91]]

3.4. Malate Dehydrogenase

Malate dehydrogenase is responsible for the reversible oxidation of malate to oxaloacetate through NAD+ to NADH reduction in the ETC, a critical step in the cellular respiration of cells. However, the role of MDH is not only limited to the ETC, as it also plays important roles in metabolic pathways, including glyoxylate bypass, amino acid synthesis, glucogenesis, and oxidation/reduction balance [[92]].

The ubiquity of MDH is related to its numerous isoforms, which have different areas of subcellular localization and co-enzyme specificity. In eukaryotic cells, there are two main isoenzymes: The mitochondrial and the cytoplasmic malate dehydrogenase. Mitochondrial malate dehydrogenase (MDH2) is critical in the citric acid cycle, as it catalyzes the reaction of malate to oxaloacetate. The other, cytosolic malate dehydrogenase (MDH1), is a key participant in the malate/aspartate shuttle and catalyzes the conversion of oxaloacetate (OAA) to malate, making transport possible. A third isoenzyme, albeit a minor one, was found in the yeast glyoxysomes, where it catalyzes the malate production from glyoxylate. For comparison, the prokaryotic Escherichia coli has only one form and is highly similar in sequence identity and tertiary structure to that of MDH2 [[92]].

Online databases, such as MalaCards, have listed diseases associated with MDH1 to include tetanus neonatorum and x-linked sideroblastic anemia with ataxia. However, a cross-reference has only presented predictability of disease occurrence, and no published literature or data can support the claim. Mutations in the MDH2 gene are related to several cancers, including uterine cancer, prostate cancer, pheochromocytoma, and other paragangliomas. MDH2 is a possible target in cancer therapeutics due to its effect on ATP production and drug sensitivity during knockdown. MDH2 was observed to be overexpressed in doxorubicin-resistant uterine cancer cells and prostate cancer cells and may contribute to drug resistance in disease models [[94]]. Its overexpression could supply more energy for P-glycoprotein in order to flush the chemotherapeutic drugs out, which may account for the shorter periods of relapse-free survival by patients with overexpressed MDH2 after chemotherapy. It is also likely that through the JNK pathway, MDH2 is able to lend docetaxel resistance in prostate cancer cells.

Recently, there has been a lot of interest in the use of potent inhibitors, such as visnagin, which holds potential cardioprotective and anticancer benefits. Researches regarding the inhibition of MDH2 as the only mechanism underlying the visnagin-induced cardioprotection are still in their early stages and require further evaluation.

Extensive research about MDH and its role in cancer is still needed, especially in the clinical aspect. In one recent study, DNA samples from 830 patients with PCCs/PGLs negative for the main PGL driver genes were analyzed. MDH2 variants of unknown significance were interpreted using an algorithm based on 20 computational predictions, enzymatic and immunofluorescence assays, and/or molecular dynamic simulation approach. The researchers identified five MDH2 variants with potential involvement in pathogenicity. Three of these variants were missense mutations and the two remaining ones were an in-frame deletion and a splice-site variant, respectively. All of the mutations were germline and are associated with noradrenergic PCCs/PGLs [[96]].

4. Future Direction of Metabolic Strategies in Combating Diseases Caused by TCA Malfunction

Although common methods such as radiotherapy and chemotherapy exert effectiveness in most patients, they sometimes pose more risks, leading to the development of cardiovascular diseases and eventual progression to heart failure. Any disruption or inefficiency in the metabolic homeostasis can undoubtedly contribute to cardiac pathologies. Such disruptions could stem from factors such as, but not limited to, the inadequate delivery of oxygen and substrates, decreased amounts of high-energy phosphate and the PCR/ATP ratio, and inefficient energy transfer or feedback [[97]].

When addressing cancer therapeutics, the effects of these strategies should also be considered, since focusing on targets such as enzymes can also prove to be detrimental to the physiological functions of normal cells and tissues. Metabolic inhibitors should minimally, if not entirely, interfere with the patient's immune system [[98]]. Nevertheless, new technology has uncovered more novel target pathways that pose less unwanted side effects to the patient.

4.1. Inhibitors and Drugs

Multiple preclinical studies have shown IDH as the recent target with the most potential for cancer drugs (Table 1). Preliminary trials indicate the selective inhibitory strength of AG-120 and AG-221 compounds in IDH1 and IDH2 mutant enzymes, respectively, by inhibiting mutant IDH activity and 2-HG accumulation, an oncometabolite [[99]]. AG-120 and another mutant IDH inhibitor, Novartis-530, were found to be the most biochemically potent inhibitors among all nine inhibitors tested in a comparative study [[101]]. In the same study, both proved to cause the highest reduction of 2-HG levels in six different cancer cell lines with IDH mutation (HT1080 fibrosarcoma, SNU1079, and RBE cholangiocarcinoma, JJ012 chondrosarcoma, U87 glioblastoma, and THP-1 AML). AG-221, on the other hand, significantly improved survival in an IDH2-mutant AML primary xenograft mouse model. Meanwhile, in preliminary phase I clinical trials in patients with advanced hematologic malignancies, the objective response rate ranged from 31% to 40%, with durable responses (>1 year) observed [[102]]. In addition, AG-881, a brain-penetrant dual IDH1/2 mutant inhibitor, is currently in phase I trial against solid tumors [[101]].

Inhibition of LDH-A, which facilitates pyruvate conversion to lactate, diminished MYC-driven tumors in xenograft models. In NSCLC mouse models, LDH-A inhibition caused the regression of established tumors without associated toxicity [[103]]. In addition, the genetic ablation of LDH-A has been observed to delay the progression of myeloid leukemia [[105]]. However, the impact of LDH-A on the adaptive immune system is yet to be explored. Lactate exhibited inhibitory action against cytotoxic T cells. Thus, blocking LDH-A activity may synergize with other immuno-inhibitors to improve host inflammatory T cell activity, which will eventually lead to the targeted tumor cells [[106]].

Meanwhile, in another clinical trial, dichloroacetate (DCA) was used in patients with lactic acidosis caused by rare inborn errors of mitochondrial metabolism. This small molecule targets pyruvate dehydrogenase kinase (PDK), an enzyme increased in different cancers due to increased activation of hypoxiainducible factor (HIF). PDK negatively regulates the pyruvate dehydrogenase complex (PDH) and blocks the oxidative decarboxylation of pyruvate to acetylCoA, which is important in leading pyruvate into the TCA cycle and away from lactate production [[107]]. Therefore, PDK inhibition by DCA causes the activation of PDH, enhanced pyruvate to acetyl CoA conversion, and decreased lactate production. More importantly, DCA is well-tolerated by patients even at doses that can affect the mitochondrial membrane potential [[108]].

4.2. Novel Approaches to TCA Targeting

However, due to the probable unwanted side effects of drug inhibitors, researchers have turned to revolutionary techniques, such as targeting microRNA (miRNA) and the use of CRISPR/Cas9 to address cancer therapy. Considering current cancer treatment regimens employing DNA-damaging drugs and/or radiation, these new approaches were conceived to be less genotoxic and cause less undesired DNA lesions in cells. This approach also outweighs the ethical limitations of mitochondrial replacement therapy. Here, we concisely discuss miRNA targeting and CRISPR/Cas9 system, both of which are promising candidates in addressing cancer therapies.

4.3. miRNA Targeting

MicroRNAs were recently brought to the spotlight due to their key roles in cancer cell metabolism, as multiple pieces of evidence have shown miRNA dysregulation in several types of cancer [[113]]. MicroRNAs are small, highly evolutionarily conserved, single-stranded, non-coding RNA molecules involved in the regulation of various gene expression. Their regulatory functions are performed through the assembly of RNA-induced silencing complex (RISC), which targets the 3′ untranslated region (UTR) of their respective mRNA [[114]]. The miRNA serves as a guide for the RISC by base-pairing with the target mRNA, and the level of complementarity between the guide and the target determines the mechanism of silencing: (1) Cleavage of target mRNA with subsequent degradation or (2) translation inhibition [[115]].

miRNAs can regulate the TCA cycle both directly and indirectly. miRNAs can downregulate subunits of pyruvate dehydrogenase (PDH), the enzyme responsible for the process that bridges glycolysis and TCA cycle—the conversion of pyruvate to acetyl CoA [[116]]. For example, miR-26a can inhibit the PDH protein X component (PDHX) [[110]], while miR-146b-5p and miR-370 can downregulate the PDHB subunit [[111]]. On the other hand, glutamine provides a major source of energy for proliferating cancer cells, and glutamine intermediates can be converted to α-KG. miR-137 targets ASC family transporter 2 (ASCT2), a glutamine transporter upregulated in different kinds of cancer. This downregulation of the transporter decreases the level of glutamine metabolism and affects cell survival in colorectal carcinoma, glioblastoma, prostate, and pancreatic cancers [[117]]. Glutamine metabolism suppression is increasingly being considered as a possible anticancer strategy. Glutaminase inhibitor CB-839 is currently undergoing a phase one clinical trial. CB-839 obstructs glutamine during the glutamate conversion process and changes the pathways of several downstream processes, such as the TCA cycle, glutathione production, and amino acid synthesis [[118]]. Epigallocatechin gallate (EGCG) reduced tumor growth in preclinical studies by interrupting the anaplerotic use of glutamine in the TCA cycle and is currently undergoing early phase one clinical trials [[119]].

In addition, miRNAs also play a pivotal role in regulating IDH. miR-183 has been shown to suppress IDH2, which causes a decrease of α-KG levels and a subsequent increase in aerobic glycolysis in glioma cells [[120]]. Meanwhile, in solid cancer tumors that rely heavily on lipid oxidation for energy source, IDH1, a cytoplasmic isoform of IDH, serves as an important contributor to lipid synthesis. This contribution can be traced back to the role of IDH in converting α-KG to isocitrate, which is subsequently converted to citrate, a precursor for the formation of monounsaturated fatty acids. miR-181a, which targets IDH1, causes a decrease of expression of genes related to lipid synthesis and increases the expression of genes involved in β-oxidation, which subsequently reduces lipid accumulation [[121]]. Moreover, miR-181a has also shown to sensitize A549 lung cancer cells to cancer drugs by stimulating Bax oligomerization and the activation of proapoptotic caspases [[122]]. The increased expression of the aforementioned miRNA was also found to increase the sensitivity of mature T cells to peptide antigens. Hence, inhibiting miR-181a in immature T cells not only reduces its sensitivity, but also impairs the positive and negative selection function of immune cells [[123]].

To take advantage of the significant role of miRNAs as post-transcriptional regulators, several miRNA-based gene therapies were developed for use against cancers. One strategy is to import exogenous tumor suppressor miRNAs, which can either inhibit the tumor cell proliferation or induce apoptosis. These exogenous miRNAs are chemically synthesized mimics of endogenous miRNAs, which are usually downregulated in tumors. These miRNA mimics can be delivered through plasmid DNA or viral vectors. Another strategy is to inhibit the function of oncogenic miRNAs using antisense oligonucleotides. In this strategy, the antagonistic oligonucleotide is complementary to the sequence of the endogenous miRNA and is chemically modified to increase its affinity with the target miRNA. This causes it to be trapped in a configuration that will either result in the inability of RISC processing or degradation of the miRNA itself [[124]].

The use of miRNA-based gene therapy against cancer is very promising (Table 2). With the advancement of genetic technology, it is highly possible that more miRNA targeting TCA enzymes will be discovered and studied. However, despite its strong logical rationale, there are still problems regarding its logistics and efficiency. miRNA-based therapies specific to TCA cycle enzymes need to be studied further in order to establish its use for TCA enzyme-related cancers.

4.4. CRISPR/Cas9 System

Clustered regularly interspaced short palindromic repeats (CRISPR) and CRISPR-associated protein 9, also known as the CRISPR/Cas9 system, was recently found as a potential therapy for cancer due to its gene-editing capability. This was discovered in prokaryotes, which possess CRISPR segments of DNA with short repetitions of base sequences. These sequences are interrupted by spacer sequences, which are remnants of viral or bacteriophage genetic codes, and thus enable recording of DNA sequences that the bacteria have been exposed to [[125]]. This then strengthens the immune function of the progeny by helping them to detect and to destroy bacteriophages once they attempt to invade the bacteria again. The spacer sequences were transcribed to RNA (crRNA), which recognizes the foreign DNA sequence. The Cas nucleases, on the other hand, mediate DNA cleavage [[126]]. For use in therapy, the spacer sequences can be genetically modified to recognize mutations in the DNA or predefined sites in the cellular genome and to facilitate cleavage.

Recently, a group of researchers, with the use of this gene-editing tool, designed sgRNAs for 88% of reported cancer mutations [[127]]. The team envisions this approach as potentially transferable to primary patient samples and for use as a therapeutic approach for personalized treatment. For example, the delivery of Cas9 and mutation-specific sgRNAs into tumor cells byoncolytic viruses could be an efficient technique for targeted therapy. Also, since specific ssgRNAs can be administered together, this strategy could be useful for combination therapy where more than two cancer mutations are targeted at the same time. However, because the method is still in its early stages, the repair mechanism after Cas9-mediated DNA cleavage is limited, resulting in sgRNA-resistant clones or off-target cleavage. In one study, this technique was also used to disrupt the CTCF motif in IDH gene in IDH wild-type gliomaspheres. The CTCF insulator protein is an important transcription factor in creating chromatin loops and boundaries that partition topological genome domains. Hypermethylation and/or disruption of binding sites of this transcription factor leads to loss of insulation between topological domains and aberrant gene activation. The aforementioned CRISPR-mediated disruption of the CTCF binding site caused an upregulation of PDGFRA, a prominent glioma oncogene, and increased cell proliferation [[128]].

In addition to the aforementioned limitations of the approach, several obstacles need to be surpassed in order to effectively use this system, such as possibilities of incomplete editing, inaccurate editing, and off-target mutations [[129]]. Such inaccuracy of the system may cause inactivation of essential genes, activation of pro-oncotic genes, or rearrangement of chromosomes. The therapy may also impose the risk of causing genetic mosaicism if it fails to affect all cells uniformly. A recent study showed that the genome editing capability of the system is affected by the tumor suppressor p53 [[130]]. p53 binds to DNA and can stimulate the transcription and activity of p21. p21 will, in turn, interact with cell division-stimulating protein (cdk2), hindering cell division. The CRISPR/Cas9 system has lower efficiency in p53 wild-type cells compared to that of knockout cells. Aside from increasing the efficiency of the CRISPR/Cas9 system, the inhibition of p53 can also decrease the selective advantage of pre-existing p53-deficient/mutant clones, a common characteristic of cancer cells. However, the inhibition of the said tumor suppressor will also increase the risk of cell vulnerability to chromosomal rearrangements and tumorigenic mutations [[130]].

The CRISPR/Cas9 system is a potentially promising cure for cancer. However, as it is still in its early stages, further research and improvements are needed to ensure its safety. Improving the balance of efficient DNA editing and suppression of potential tumorigenic effects is also important. Considering the ease of its use, its application to other TCA enzymes opens another area for medical research.

5. Conclusions

Understanding the mechanisms involving mitochondria is important to further develop strategies in treating mitochondria-associated diseases and dysfunction. The majority of studies to date have focused more on OXPHOS, which is another process that occurs in the mitochondrion and connects the mitochrondion to cancer physiology. However, the TCA cycle is also now being recognized as a key player in certain cancers which involve enzyme dysfunction. This review presented different studies regarding these mitochondrial TCA enzymes and cited diseases where they play a pivotal role. In addition, we also discussed available and prospective treatments, such as the drugs mentioned in the previous section. Last, this review was able to explore two novel approaches which are both promising strategies for cancer treatment: CRISPR/Cas9 and microRNA. However, both strategies require further research to ensure their specificity and efficiency.

Figures and Tables

Graph: Figure 1 Mutations in the tricarboxylic acid (TCA) cycle enzymes may result in various kinds of cancers. These diseases may possibly be treated through pharmacological and genetic therapeutic approaches. IDH, isocitrate dehydrogenase; SDH, succinate dehydrogenase; FH, fumarate hydratase; MDH, malate dehydrogenase.

Graph: Figure 2 Schematic presentation of the difference in cellular pathways of wild-type and mutated IDH1/2 enzymes during reverse reductive carboxylation reaction. IDH1/2 enzymes catalyzes both the forward and reverse conversion of isocitrate to αKG. Mutations in IDH1/2 cause elevated levels of R-2HG (D-2HG), which is a pro-oncogenic factor. αKG, α-ketoglutarate; ICT, isocitrate; CT, citrate; CS, citrate synthase; FH, fumarate hydratase; FAs, fatty acids. Adapted from Al-Khallaf H. (2017) [[26]].

Graph: Figure 3 Schematic diagram of tumorigenesis in succinate dehydrogenase (SDH) and fumarate hydratase (FH). SDH and FH deficiency causes accumulation of succinate and fumarate, respectively, inside the mitochondria. These will be subsequently transported into the cytosol. High levels of succinate and fumarate can inhibit prolyl hydroxylases (PDH1-3), which plays a role in the degradation of HIF-1α under normoxic conditions. HIF-1α, when stabilized, induces transcription of nuclear genes involved in tumor suppression. αKG, α-ketoglutarate; ICT, isocitrate; CT, citrate; SCoA, Succinyl-coA; HIF, hypoxia-inducible factor Adapted from Zanssen S, Schon EA (2005) [[49]] and Shuch, B., Linehan, W.M., & Srinivasan, R. (2013) [[50]].

Table 1 Drugs and inhibitors and their respective TCA or TCA-related enzyme targets.

Drug/InhibitorTargetRole of TargetAction of Drug to TargetSample TypeReference
AG-120 (Ivosidenib)IDH1Catalyze conversion of isocitrate to α-ketoglutarateinhibitClinical trial: glioma, adcanced hematologic malignancy[87-88]
AG-221IDH2Catalyze conversion of isocitrate to α-ketoglutarateinhibitClinical trial: acute myeloid leukemia and myelodysplastic syndrome patients[86]
Novartis-530IDH1Catalyze conversion of isocitrate to α-ketoglutarateinhibitCancer cell lines with somatic IDH1 mutation[88]
FX 11LDH-AForward and reverse conversion of pyruvate to lactateinhibithuman lymphoma and pancreatic cancer xenografts[91]
Dichloroacetate (DCA)PDKPhosphorylation and inhibition of PDCinhibitHuman lung carcinoma cell[95-96]

Table 2 miRNAs and their respective TCA or TCA-related enzyme targets.

miRNA NameTargetRole of TargetAction of miRNA to TargetSample TypeReference
miR-26aPDHXCatalyzes conversion of pyruvate to acetyl coAinhibitColorectal cancer cell lines[110]
miR-146b-5pPDHBConversion of glucose-derived pyruvate to acetyl coAinhibitHuman colorectal cancer tissue samples, colorectal cancer cell lines[111]
miR-370PDHBConversion of glucose-derived pyruvate to acetyl coAinhibitHuman melanoma tissue samples, human melanoma cell line[112]
miR-137ASCT2Transport of glutamineinhibitHuman neuroblastoma cell line[107]
miR-183IDH2Catalyze conversion of isocitrate to α-ketoglutarateinhibitGlioblastoma cell lines[108]
miR-181aIDH1Catalyze conversion of isocitrate to α-ketoglutarateinhibitTail-tip fibroblast, mouse embryonic fibroblast[109]
Human lung cancer cell line, human colon cancer cell line, human cervical cancer cell line[110]
Mouse T-cells[111]

Author Contributions

Conceptualization of this review paper was done by J.M. and J.H.; writing and the original draft was by J.M., J.F., B.N., A.T.T.N., N.P., J.H.; review, and editing by J.M., A.H.K., N.P., and J.H.; visualization by J.F.; supervised byN.P., J.H.; and funding acquisition, J.H.

Funding

This work was supported by the Priority Research Centers Program (2010-0020224). This work was also supported by the National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (2018R1A2A3074998).

Conflicts of Interest

The authors declare no conflict of interest.

References 1 Vyas S., Zaganjor E., Haigis M.C. Mitochondria and Cancer. Cell. 2016; 166: 555-566. 10.1016/j.cell.2016.07.002. 27471965 2 Calvo S.E., Mootha V.K. The mitochondrial proteome and human disease. Annu. Rev. Genom. Hum. Genet. 2010; 11: 25-44. 10.1146/annurev-genom-082509-141720. 20690818 3 Taylor R.W., Turnbull D.M. Mitochondrial DNA mutations in human disease. Nat. Rev. Genet. 2005; 6: 389-402. 10.1038/nrg1606. 15861210 4 Lu P., Bruno B.J., Rabenau M., Lim C.S. Delivery of drugs and macromolecules to the mitochondria for cancer therapy. J. Control. Release. 2016; 240: 38-51. 10.1016/j.jconrel.2015.10.023. 26482081 5 Yang Y., Karakhanova S., Hartwig W., D'Haese J.G., Philippov P.P., Werner J., Bazhin A.V. Mitochondria and Mitochondrial ROS in Cancer: Novel Targets for Anticancer Therapy. J. Cell. Physiol. 2016; 231: 2570-2581. 10.1002/jcp.25349. 26895995 6 Beutner G., Alavian K.N., Jonas E.A., Porter G.A. Jr. The Mitochondrial Permeability Transition Pore and ATP Synthase. Handb. Exp. Pharmacol. 2016. 10.1007/164_2016_5 7 Rogers G.W., Nadanaciva S., Swiss R., Divakaruni A.S., Will Y. Assessment of fatty acid beta oxidation in cells and isolated mitochondria. Curr. Protoc. Toxicol. 2014; 60: 25.3.1-25.3.19. 10.1002/0471140856.tx2503s60 8 Mergenthaler P., Lindauer U., Dienel G.A., Meisel A. Sugar for the brain: The role of glucose in physiological and pathological brain function. Trends Neurosci. 2013; 36: 587-597. 10.1016/j.tins.2013.07.001 9 Jensen J., Rustad P.I., Kolnes A.J., Lai Y.C. The role of skeletal muscle glycogen breakdown for regulation of insulin sensitivity by exercise. Front. Physiol. 2011; 2: 112. 10.3389/fphys.2011.00112 Rui L. Energy metabolism in the liver. Compr. Physiol. 2014; 4: 177-197. 10.1002/cphy.c130024 Nguyen P., Leray V., Diez M., Serisier S., Le Bloc'h J., Siliart B., Dumon H. Liver lipid metabolism. J. Anim. Physiol. Anim. Nutr. 2008; 92: 272-283. 10.1111/j.1439-0396.2007.00752.x. 18477307 Owen O.E., Kalhan S.C., Hanson R.W. The key role of anaplerosis and cataplerosis for citric acid cycle function. J. Biol. Chem. 2002; 277: 30409-30412. 10.1074/jbc.R200006200. 12087111 Gibala M.J., Young M.E., Taegtmeyer H. Anaplerosis of the citric acid cycle: Role in energy metabolism of heart and skeletal muscle. Acta Physiol. Scand. 2000; 168: 657-665. 10.1046/j.1365-201x.2000.00717.x. 10759602 Ahn C.S., Metallo C.M. Mitochondria as biosynthetic factories for cancer proliferation. Cancer Metab. 2015; 3: 1. 10.1186/s40170-015-0128-2. 25621173 Cheng T., Sudderth J., Yang C., Mullen A.R., Jin E.S., Mates J.M., DeBerardinis R.J. Pyruvate carboxylase is required for glutamine-independent growth of tumor cells. Proc. Natl. Acad. Sci. USA. 2011; 108: 8674-8679. 10.1073/pnas.1016627108. 21555572 Kolwicz S.C. Jr., Purohit S., Tian R. Cardiac metabolism and its interactions with contraction, growth, and survival of cardiomyocytes. Circ. Res. 2013; 113: 603-616. 10.1161/CIRCRESAHA.113.302095. 23948585 Smith A.C., Robinson A.J. A metabolic model of the mitochondrion and its use in modelling diseases of the tricarboxylic acid cycle. BMC Syst. Biol. 2011; 5102. 10.1186/1752-0509-5-102 Seyfried T.N., Flores R.E., Poff A.M., D'Agostino D.P. Cancer as a metabolic disease: Implications for novel therapeutics. Carcinogenesis. 2014; 35: 515-527. 10.1093/carcin/bgt480 Seyfried T.N. Cancer as a mitochondrial metabolic disease. Front. Cell Dev. Biol. 2015; 3: 43. 10.3389/fcell.2015.00043 Warburg O. On the origin of cancer cells. Science. 1956; 123: 309-314. 10.1126/science.123.3191.309 Pavlova N.N., Thompson C.B. The Emerging Hallmarks of Cancer Metabolism. Cell Metab. 2016; 23: 27-47. 10.1016/j.cmet.2015.12.006. 26771115 Courtnay R., Ngo D.C., Malik N., Ververis K., Tortorella S.M., Karagiannis T.C. Cancer metabolism and the Warburg effect: The role of HIF-1 and PI3K. Mol. Biol. Rep. 2015; 42: 841-851. 10.1007/s11033-015-3858-x. 25689954 Payne S.R., Kemp C.J. Tumor suppressor genetics. Carcinogenesis. 2005; 26: 2031-2045. 10.1093/carcin/bgi223. 16150895 Chen J.Q., Russo J. Dysregulation of glucose transport, glycolysis, TCA cycle and glutaminolysis by oncogenes and tumor suppressors in cancer cells. Biochim. Biophys. Acta. 2012; 1826: 370-384. 10.1016/j.bbcan.2012.06.004. 22750268 Laurenti G., Tennant D.A. Isocitrate dehydrogenase (IDH), succinate dehydrogenase (SDH), fumarate hydratase (FH): Three players for one phenotype in cancer?. Biochem. Soc. Trans. 2016; 44: 1111-1116. 10.1042/BST20160099 Al-Khallaf H. Isocitrate dehydrogenases in physiology and cancer: Biochemical and molecular insight. Cell Biosci. 2017; 7: 37. 10.1186/s13578-017-0165-3 Stoddard B.L., Dean A., Koshland D.E. Jr. Structure of isocitrate dehydrogenase with isocitrate, nicotinamide adenine dinucleotide phosphate, and calcium at 2.5-A resolution: A pseudo-Michaelis ternary complex. Biochemistry. 1993; 32: 9310-9316. 10.1021/bi00087a008 Mondesir J., Willekens C., Touat M., de Botton S. IDH1 and IDH2 mutations as novel therapeutic targets: Current perspectives. J. Blood Med. 2016; 7: 171-180. 10.2147/JBM.S70716 Cairns R.A., Iqbal J., Lemonnier F., Kucuk C., de Leval L., Jais J.P., Parrens M., Martin A., Xerri L., Brousset P. IDH2 mutations are frequent in angioimmunoblastic T-cell lymphoma. Blood. 2012; 119: 1901-1903. 10.1182/blood-2011-11-391748 Kang M.R., Kim M.S., Oh J.E., Kim Y.R., Song S.Y., Seo S.I., Lee J.Y., Yoo N.J., Lee S.H. Mutational analysis of IDH1 codon 132 in glioblastomas and other common cancers. Int. J. Cancer. 2009; 125: 353-355. 10.1002/ijc.24379 Yen K.E., Bittinger M.A., Su S.M., Fantin V.R. Cancer-associated IDH mutations: Biomarker and therapeutic opportunities. Oncogene. 2010; 29: 6409-6417. 10.1038/onc.2010.444 Abbas S., Lugthart S., Kavelaars F.G., Schelen A., Koenders J.E., Zeilemaker A., van Putten W.J., Rijneveld A.W., Lowenberg B., Valk P.J. Acquired mutations in the genes encoding IDH1 and IDH2 both are recurrent aberrations in acute myeloid leukemia: Prevalence and prognostic value. Blood. 2010; 116: 2122-2126. 10.1182/blood-2009-11-250878. 20538800 Paschka P., Schlenk R.F., Gaidzik V.I., Habdank M., Kronke J., Bullinger L., Spath D., Kayser S., Zucknick M., Gotze K. IDH1 and IDH2 mutations are frequent genetic alterations in acute myeloid leukemia and confer adverse prognosis in cytogenetically normal acute myeloid leukemia with NPM1 mutation without FLT3 internal tandem duplication. J. Clin. Oncol. 2010; 28: 3636-3643. 10.1200/JCO.2010.28.3762. 20567020 Xu W., Yang H., Liu Y., Yang Y., Wang P., Kim S.H., Ito S., Yang C., Xiao M.T., Liu L.X. Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of alpha-ketoglutarate-dependent dioxygenases. Cancer Cell. 2011; 19: 17-30. 10.1016/j.ccr.2010.12.014. 21251613 Carbonneau M., Gagné L.M., Lalonde M.E., Germain M.A., Motorina A., Guiot M.C., Secco B., Vincent E.E., Tumber A., Hulea L. The oncometabolite 2-hydroxyglutarate activates the mTOR signalling pathway. Nat. Commun. 2016; 7: 12700. 10.1038/ncomms12700 Populo H., Lopes J.M., Soares P. The mTOR signalling pathway in human cancer. Int. J. Mol. Sci. 2012; 13: 1886-1918. 10.3390/ijms13021886 Whetstine J.R., Nottke A., Lan F., Huarte M., Smolikov S., Chen Z., Spooner E., Li E., Zhang G., Colaiacovo M. Reversal of histone lysine trimethylation by the JMJD2 family of histone demethylases. Cell. 2006; 125: 467-481. 10.1016/j.cell.2006.03.028 Li T., Cox C.D., Ozer B.H., Nguyen N.T., Nguyen H.N., Lai T.J., Li S., Liu F., Kornblum H.I., Liau L.M. D-2-Hydroxyglutarate Is Necessary and Sufficient for Isocitrate Dehydrogenase 1 Mutant-Induced MIR148A Promoter Methylation. Mol. Cancer Res. 2018; 16: 947-960. 10.1158/1541-7786.MCR-17-0367 Dang L., White D.W., Gross S., Bennett B.D., Bittinger M.A., Driggers E.M., Fantin V.R., Jang H.G., Jin S., Keenan M.C. Cancer-associated IDH1 mutations produce 2-hydroxyglutarate. Nature. 2009; 462: 739-744. 10.1038/nature08617 Losman J.A., Kaelin W.G. Jr. What a difference a hydroxyl makes: Mutant IDH, (R)-2-hydroxyglutarate, and cancer. Genes Dev. 2013; 27: 836-852. 10.1101/gad.217406.113 Mardis E.R., Ding L., Dooling D.J., Larson D.E., McLellan M.D., Chen K., Koboldt D.C., Fulton R.S., Delehaunty K.D., McGrath S.D. Recurring mutations found by sequencing an acute myeloid leukemia genome. N. Engl. J. Med. 2009; 361: 1058-1066. 10.1056/NEJMoa0903840. 19657110 Im A.P., Sehgal A.R., Carroll M.P., Smith B.D., Tefferi A., Johnson D.E., Boyiadzis M. DNMT3A and IDH mutations in acute myeloid leukemia and other myeloid malignancies: Associations with prognosis and potential treatment strategies. Leukemia. 2014; 28: 1774-1783. 10.1038/leu.2014.124. 24699305 Molenaar R.J., Thota S., Nagata Y., Patel B., Clemente M., Przychodzen B., Hirsh C., Viny A.D., Hosano N., Bleeker F.E. Clinical and biological implications of ancestral and non-ancestral IDH1 and IDH2 mutations in myeloid neoplasms. Leukemia. 2015; 29: 2134-2142. 10.1038/leu.2015.91. 25836588 Patel J.P., Gonen M., Figueroa M.E., Fernandez H., Sun Z., Racevskis J., Van Vlierberghe P., Dolgalev I., Thomas S., Aminova O. Prognostic relevance of integrated genetic profiling in acute myeloid leukemia. N. Engl. J. Med. 2012; 366: 1079-1089. 10.1056/NEJMoa1112304. 22417203 Abdel-Wahab O., Patel J., Levine R.L. Clinical implications of novel mutations in epigenetic modifiers in AML. Hematol. Oncol. Clin. N. Am. 2011; 25: 1119-1133. 10.1016/j.hoc.2011.09.013. 22093580 Boissel N., Nibourel O., Renneville A., Huchette P., Dombret H., Preudhomme C. Differential prognosis impact of IDH2 mutations in cytogenetically normal acute myeloid leukemia. Blood. 2011; 117: 3696-3697. 10.1182/blood-2010-11-320937 Green C.L., Evans C.M., Zhao L., Hills R.K., Burnett A.K., Linch D.C., Gale R.E. The prognostic significance of IDH2 mutations in AML depends on the location of the mutation. Blood. 2011; 118: 409-412. 10.1182/blood-2010-12-322479 Hamadou W.S., Bourdon V., Letard S., Brenet F., Laarif S., Besbes S., Paci A., David M., Penard-Lacronique V., Youssef Y.B. Familial hematological malignancies: New IDH2 mutation. Ann. Hematol. 2016. 10.1007/s00277-016-2813-9 Zanssen S., Schon E.A. Mitochondrial DNA mutations in cancer. PLoS Med. 2005; 2e401. 10.1371/journal.pmed.0020401 Shuch B., Linehan W.M., Srinivasan R. Aerobic glycolysis: A novel target in kidney cancer. Expert Rev. Anticancer Ther. 2013; 13: 711-719. 10.1586/era.13.57 Baysal B.E., Ferrell R.E., Willett-Brozick J.E., Lawrence E.C., Myssiorek D., Bosch A., van der Mey A., Taschner P.E., Rubinstein W.S., Myers E.N. Mutations in SDHD, a mitochondrial complex II gene, in hereditary paraganglioma. Science. 2000; 287: 848-851. 10.1126/science.287.5454.848. 10657297 Selak M.A., Armour S.M., MacKenzie E.D., Boulahbel H., Watson D.G., Mansfield K.D., Pan Y., Simon M.C., Thompson C.B., Gottlieb E. Succinate links TCA cycle dysfunction to oncogenesis by inhibiting HIF-alpha prolyl hydroxylase. Cancer Cell. 2005; 7: 77-85. 10.1016/j.ccr.2004.11.022. 15652751 Ryan D.G., Murphy M.P., Frezza C., Prag H.A., Chouchani E.T., O'Neill L.A., Mills E.L. Coupling Krebs cycle metabolites to signalling in immunity and cancer. Nat. Metab. 2019; 1: 16-33. 10.1038/s42255-018-0014-7. 31032474 Niemann S., Muller U. Mutations in SDHC cause autosomal dominant paraganglioma, type 3. Nat. Genet. 2000; 26: 268-270. 10.1038/81551 Astuti D., Latif F., Dallol A., Dahia P.L., Douglas F., George E., Skoldberg F., Husebye E.S., Eng C., Maher E.R. Gene mutations in the succinate dehydrogenase subunit SDHB cause susceptibility to familial pheochromocytoma and to familial paraganglioma. Am. J. Hum. Genet. 2001; 69: 49-54. 10.1086/321282 Bardella C., Pollard P.J., Tomlinson I. SDH mutations in cancer. Biochim. Biophys. Acta. 2011; 1807: 1432-1443. 10.1016/j.bbabio.2011.07.003 Bourgeron T., Rustin P., Chretien D., Birch-Machin M., Bourgeois M., Viegas-Pequignot E., Munnich A., Rotig A. Mutation of a nuclear succinate dehydrogenase gene results in mitochondrial respiratory chain deficiency. Nat. Genet. 1995; 11: 144-149. 10.1038/ng1095-144 Parfait B., Chretien D., Rotig A., Marsac C., Munnich A., Rustin P. Compound heterozygous mutations in the flavoprotein gene of the respiratory chain complex II in a patient with Leigh syndrome. Hum. Genet. 2000; 106: 236-243. 10.1007/s004390051033 Horvath R., Abicht A., Holinski-Feder E., Laner A., Gempel K., Prokisch H., Lochmuller H., Klopstock T., Jaksch M. Leigh syndrome caused by mutations in the flavoprotein (Fp) subunit of succinate dehydrogenase (SDHA). J. Neurol. Neurosurg. Psychiatry. 2006; 77: 74-76. 10.1136/jnnp.2005.067041 Birch-Machin M.A., Taylor R.W., Cochran B., Ackrell B.A., Turnbull D.M. Late-onset optic atrophy, ataxia, and myopathy associated with a mutation of a complex II gene. Ann. Neurol. 2000; 48: 330-335. 10.1002/1531-8249(200009)48:3<330::AID-ANA7>3.0.CO;2-A Van Coster R., Seneca S., Smet J., Van Hecke R., Gerlo E., Devreese B., Van Beeumen J., Leroy J.G., De Meirleir L., Lissens W. Homozygous Gly555Glu mutation in the nuclear-encoded 70 kDa flavoprotein gene causes instability of the respiratory chain complex II. Am. J. Med. Genet. A. 2003; 120: 13-18. 10.1002/ajmg.a.10202. 12794685 Lin G., Doyle L.A. An update on the application of newly described immunohistochemical markers in soft tissue pathology. Arch. Pathol. Lab. Med. 2015; 139: 106-121. 10.5858/arpa.2014-0488-RA. 25549147 Neumann H.P., Pawlu C., Peczkowska M., Bausch B., McWhinney S.R., Muresan M., Buchta M., Franke G., Klisch J., Bley T.A. Distinct clinical features of paraganglioma syndromes associated with SDHB and SDHD gene mutations. JAMA. 2004; 292: 943-951. 10.1001/jama.292.8.943. 15328326 Timmers H.J., Kozupa A., Eisenhofer G., Raygada M., Adams K.T., Solis D., Lenders J.W., Pacak K. Clinical presentations, biochemical phenotypes, and genotype-phenotype correlations in patients with succinate dehydrogenase subunit B-associated pheochromocytomas and paragangliomas. J. Clin. Endocrinol. Metab. 2007; 92: 779-786. 10.1210/jc.2006-2315. 17200167 Benn D.E., Gimenez-Roqueplo A.P., Reilly J.R., Bertherat J., Burgess J., Byth K., Croxson M., Dahia P.L., Elston M., Gimm O. Clinical presentation and penetrance of pheochromocytoma/paraganglioma syndromes. J. Clin. Endocrinol. Metab. 2006; 91: 827-836. 10.1210/jc.2005-1862. 16317055 Young A.L., Baysal B.E., Deb A., Young W.F. Jr. Familial malignant catecholamine-secreting paraganglioma with prolonged survival associated with mutation in the succinate dehydrogenase B gene. J. Clin. Endocrinol. Metab. 2002; 87: 4101-4105. 10.1210/jc.2002-020312 Vanharanta S., Buchta M., McWhinney S.R., Virta S.K., Peczkowska M., Morrison C.D., Lehtonen R., Januszewicz A., Jarvinen H., Juhola M. Early-onset renal cell carcinoma as a novel extraparaganglial component of SDHB-associated heritable paraganglioma. Am. J. Hum. Genet. 2004; 74: 153-159. 10.1086/381054 Baysal B.E. A recurrent stop-codon mutation in succinate dehydrogenase subunit B gene in normal peripheral blood and childhood T-cell acute leukemia. PLoS ONE. 2007; 2e436. 10.1371/journal.pone.0000436 Fishbein L., Leshchiner I., Walter V., Danilova L., Robertson A.G., Johnson A.R., Lichtenberg T.M., Murray B.A., Ghayee H.K., Else T. Comprehensive Molecular Characterization of Pheochromocytoma and Paraganglioma. Cancer Cell. 2017; 31: 181-193. 10.1016/j.ccell.2017.01.001 Burnichon N., Rohmer V., Amar L., Herman P., Leboulleux S., Darrouzet V., Niccoli P., Gaillard D., Chabrier G., Chabolle F. The succinate dehydrogenase genetic testing in a large prospective series of patients with paragangliomas. J. Clin. Endocrinol. Metab. 2009; 94: 2817-2827. 10.1210/jc.2008-2504 Mannelli M., Ercolino T., Giache V., Simi L., Cirami C., Parenti G. Genetic screening for pheochromocytoma: Should SDHC gene analysis be included?. J. Med. Genet. 2007; 44: 586-587. 10.1136/jmg.2007.051045. 17557926 Mannelli M., Castellano M., Schiavi F., Filetti S., Giacche M., Mori L., Pignataro V., Bernini G., Giache V., Bacca A. Clinically guided genetic screening in a large cohort of italian patients with pheochromocytomas and/or functional or nonfunctional paragangliomas. J. Clin. Endocrinol. Metab. 2009; 94: 1541-1547. 10.1210/jc.2008-2419. 19223516 Schiavi F., Boedeker C.C., Bausch B., Peczkowska M., Gomez C.F., Strassburg T., Pawlu C., Buchta M., Salzmann M., Hoffmann M.M. Predictors and prevalence of paraganglioma syndrome associated with mutations of the SDHC gene. JAMA. 2005; 294: 2057-2063. 10.1001/jama.294.16.2057. 16249420 Ni Y., Seballos S., Ganapathi S., Gurin D., Fletcher B., Ngeow J., Nagy R., Kloos R.T., Ringel M.D., LaFramboise T. Germline and somatic SDHx alterations in apparently sporadic differentiated thyroid cancer. Endocr. Relat. Cancer. 2015; 22: 121-130. 10.1530/ERC-14-0537. 25694510 Ricketts C.J., Forman J.R., Rattenberry E., Bradshaw N., Lalloo F., Izatt L., Cole T.R., Armstrong R., Kumar V.K., Morrison P.J. Tumor risks and genotype-phenotype-proteotype analysis in 358 patients with germline mutations in SDHB and SDHD. Hum. Mutat. 2010; 31: 41-51. 10.1002/humu.21136. 19802898 Bennedbaek M., Rossing M., Rasmussen A.K., Gerdes A.M., Skytte A.B., Jensen U.B., Nielsen F.C., Hansen T.V. Identification of eight novel SDHB, SDHC, SDHD germline variants in Danish pheochromocytoma/paraganglioma patients. Hered. Cancer Clin. Pract. 2016; 14: 13. 10.1186/s13053-016-0053-6. 27279923 Burnichon N., Briere J.J., Libe R., Vescovo L., Riviere J., Tissier F., Jouanno E., Jeunemaitre X., Benit P., Tzagoloff A. SDHA is a tumor suppressor gene causing paraganglioma. Hum. Mol. Genet. 2010; 19: 3011-3020. 10.1093/hmg/ddq206 Bayley J.P., Kunst H.P., Cascon A., Sampietro M.L., Gaal J., Korpershoek E., Hinojar-Gutierrez A., Timmers H.J., Hoefsloot L.H., Hermsen M.A. SDHAF2 mutations in familial and sporadic paraganglioma and phaeochromocytoma. Lancet Oncol. 2010; 11: 366-372. 10.1016/S1470-2045(10)70007-3 Hao H.X., Khalimonchuk O., Schraders M., Dephoure N., Bayley J.P., Kunst H., Devilee P., Cremers C.W., Schiffman J.D., Bentz B.G. SDH5, a gene required for flavination of succinate dehydrogenase, is mutated in paraganglioma. Science. 2009; 325: 1139-1142. 10.1126/science.1175689 Zhang D.D., Hannink M. Distinct cysteine residues in Keap1 are required for Keap1-dependent ubiquitination of Nrf2 and for stabilization of Nrf2 by chemopreventive agents and oxidative stress. Mol. Cell. Biol. 2003; 23: 8137-8151. 10.1128/MCB.23.22.8137-8151.2003 Ma Q. Role of nrf2 in oxidative stress and toxicity. Annu. Rev. Pharmacol. Toxicol. 2013; 53: 401-426. 10.1146/annurev-pharmtox-011112-140320. 23294312 Tyrakis P.A., Yurkovich M.E., Sciacovelli M., Papachristou E.K., Bridges H.R., Gaude E., Schreiner A., D'Santos C., Hirst J., Hernandez-Fernaud J. Fumarate Hydratase Loss Causes Combined Respiratory Chain Defects. Cell Rep. 2017; 21: 1036-1047. 10.1016/j.celrep.2017.09.092. 29069586 Launonen V., Vierimaa O., Kiuru M., Isola J., Roth S., Pukkala E., Sistonen P., Herva R., Aaltonen L.A. Inherited susceptibility to uterine leiomyomas and renal cell cancer. Proc. Natl. Acad. Sci. USA. 2001; 98: 3387-3392. 10.1073/pnas.051633798. 11248088 Tomlinson I.P., Alam N.A., Rowan A.J., Barclay E., Jaeger E.E., Kelsell D., Leigh I., Gorman P., Lamlum H., Rahman S. Germline mutations in FH predispose to dominantly inherited uterine fibroids, skin leiomyomata and papillary renal cell cancer. Nat. Genet. 2002; 30: 406-410. 10.1038/ng849 Barker K.T., Bevan S., Wang R., Lu Y.J., Flanagan A.M., Bridge J.A., Fisher C., Finlayson C.J., Shipley J., Houlston R.S. Low frequency of somatic mutations in the FH/multiple cutaneous leiomyomatosis gene in sporadic leiomyosarcomas and uterine leiomyomas. Br. J. Cancer. 2002; 87: 446-448. 10.1038/sj.bjc.6600502 Letouze E., Martinelli C., Loriot C., Burnichon N., Abermil N., Ottolenghi C., Janin M., Menara M., Nguyen A.T., Benit P. SDH mutations establish a hypermethylator phenotype in paraganglioma. Cancer Cell. 2013; 23: 739-752. 10.1016/j.ccr.2013.04.018 Smit D.L., Mensenkamp A.R., Badeloe S., Breuning M.H., Simon M.E., van Spaendonck K.Y., Aalfs C.M., Post J.G., Shanley S., Krapels I.P. Hereditary leiomyomatosis and renal cell cancer in families referred for fumarate hydratase germline mutation analysis. Clin. Genet. 2011; 79: 49-59. 10.1111/j.1399-0004.2010.01486.x Schmidt L.S., Linehan W.M. Hereditary leiomyomatosis and renal cell carcinoma. Int. J. Nephrol. Renovasc. Dis. 2014; 7: 253-260. 10.2147/IJNRD.S42097 Zhao S., Lin Y., Xu W., Jiang W., Zha Z., Wang P., Yu W., Li Z., Gong L., Peng Y. Glioma-derived mutations in IDH1 dominantly inhibit IDH1 catalytic activity and induce HIF-1alpha. Science. 2009; 324: 261-265. 10.1126/science.1170944 Sudarshan S., Shanmugasundaram K., Naylor S.L., Lin S., Livi C.B., O'Neill C.F., Parekh D.J., Yeh I.T., Sun L.Z., Block K. Reduced expression of fumarate hydratase in clear cell renal cancer mediates HIF-2alpha accumulation and promotes migration and invasion. PLoS ONE. 2011; 6e21037. 10.1371/journal.pone.0021037 Tong W.H., Sourbier C., Kovtunovych G., Jeong S.Y., Vira M., Ghosh M., Romero V.V., Sougrat R., Vaulont S., Viollet B. The glycolytic shift in fumarate-hydratase-deficient kidney cancer lowers AMPK levels, increases anabolic propensities and lowers cellular iron levels. Cancer Cell. 2011; 20: 315-327. 10.1016/j.ccr.2011.07.018. 21907923 Musrati R.A., Kollarova M., Mernik N., Mikulasova D. Malate dehydrogenase: Distribution, function and properties. Gen. Physiol. Biophys. 1998; 17: 193-210. 9834842 Minarik P., Tomaskova N., Kollarova M., Antalik M. Malate dehydrogenases--structure and function. Gen. Physiol. Biophys. 2002; 21: 257-265. 12537350 Lo Y.W., Lin S.T., Chang S.J., Chan C.H., Lyu K.W., Chang J.F., May E.W., Lin D.Y., Chou H.C., Chan H.L. Mitochondrial proteomics with siRNA knockdown to reveal ACAT1 and MDH2 in the development of doxorubicin-resistant uterine cancer. J. Cell. Mol. Med. 2015; 19: 744-759. 10.1111/jcmm.12388 Liu Q., Harvey C.T., Geng H., Xue C., Chen V., Beer T.M., Qian D.Z. Malate dehydrogenase 2 confers docetaxel resistance via regulations of JNK signaling and oxidative metabolism. Prostate. 2013; 73: 1028-1037. 10.1002/pros.22650 Calsina B., Curras-Freixes M., Buffet A., Pons T., Contreras L., Leton R., Comino-Mendez I., Remacha L., Calatayud M., Obispo B. Role of MDH2 pathogenic variant in pheochromocytoma and paraganglioma patients. Genet. Med. 2018; 20: 1652-1662. 10.1038/s41436-018-0068-7 Ventura-Clapier R., Garnier A., Veksler V. Energy metabolism in heart failure. J. Physiol. 2004; 555: 1-13. 10.1113/jphysiol.2003.055095 Erez A., DeBerardinis R.J. Metabolic dysregulation in monogenic disorders and cancer—Finding method in madness. Nat. Rev. Cancer. 2015; 15: 440-448. 10.1038/nrc3949 Chen J., Yang J., Cao P. The Evolving Landscape in the Development of Isocitrate Dehydrogenase Mutant Inhibitors. Mini Rev. Med. Chem. 2016; 16: 1344-1358. 10.2174/1389557516666160609085520 Caino M.C., Altieri D.C. Molecular Pathways: Mitochondrial Reprogramming in Tumor Progression and Therapy. Clin. Cancer Res. 2016; 22: 540-545. 10.1158/1078-0432.CCR-15-0460 Urban D.J., Martinez N.J., Davis M.I., Brimacombe K.R., Cheff D.M., Lee T.D., Henderson M.J., Titus S.A., Pragani R., Rohde J.M. Assessing inhibitors of mutant isocitrate dehydrogenase using a suite of pre-clinical discovery assays. Sci. Rep. 2017; 7: 12758. 10.1038/s41598-017-12630-x. 28986582 Fujii T., Khawaja M.R., DiNardo C.D., Atkins J.T., Janku F. Targeting isocitrate dehydrogenase (IDH) in cancer. Discov. Med. 2016; 21: 373-380. 27355333 Fantin V.R., St-Pierre J., Leder P. Attenuation of LDH-A expression uncovers a link between glycolysis, mitochondrial physiology, and tumor maintenance. Cancer Cell. 2006; 9: 425-434. 10.1016/j.ccr.2006.04.023 Le A., Cooper C.R., Gouw A.M., Dinavahi R., Maitra A., Deck L.M., Royer R.E., Vander Jagt D.L., Semenza G.L., Dang C.V. Inhibition of lactate dehydrogenase A induces oxidative stress and inhibits tumor progression. Proc. Natl. Acad. Sci. USA. 2010; 107: 2037-2042. 10.1073/pnas.0914433107 DeBerardinis R.J., Chandel N.S. Fundamentals of cancer metabolism. Sci. Adv. 2016; 2: e1600200. 10.1126/sciadv.1600200. 27386546 Haas R., Smith J., Rocher-Ros V., Nadkarni S., Montero-Melendez T., D'Acquisto F., Bland E.J., Bombardieri M., Pitzalis C., Perretti M. Lactate Regulates Metabolic and Pro-inflammatory Circuits in Control of T Cell Migration and Effector Functions. PLoS Biol. 2015; 13e1002202. 10.1371/journal.pbio.1002202. 26181372 Bonnet S., Archer S.L., Allalunis-Turner J., Haromy A., Beaulieu C., Thompson R., Lee C.T., Lopaschuk G.D., Puttagunta L., Bonnet S. A mitochondria-K+ channel axis is suppressed in cancer and its normalization promotes apoptosis and inhibits cancer growth. Cancer Cell. 2007; 11: 37-51. 10.1016/j.ccr.2006.10.020 Michelakis E.D., Sutendra G., Dromparis P., Webster L., Haromy A., Niven E., Maguire C., Gammer T.L., Mackey J.R., Fulton D. Metabolic modulation of glioblastoma with dichloroacetate. Sci. Transl. Med. 2010; 2: 31ra34. 10.1126/scitranslmed.3000677 Vander Heiden M.G. Targeting cancer metabolism: A therapeutic window opens. Nat. Rev. Drug Discov. 2011; 10: 671-684. 10.1038/nrd3504 Chen B., Liu Y., Jin X., Lu W., Liu J., Xia Z., Yuan Q., Zhao X., Xu N., Liang S. MicroRNA-26a regulates glucose metabolism by direct targeting PDHX in colorectal cancer cells. BMC Cancer. 2014; 14443. 10.1186/1471-2407-14-443 Zhu Y., Wu G., Yan W., Zhan H., Sun P. miR-146b-5p regulates cell growth, invasion, and metabolism by targeting PDHB in colorectal cancer. Am. J. Cancer Res. 2017; 7: 1136-1150. 28560062 Wei S., Ma W. MiR-370 functions as oncogene in melanoma by direct targeting pyruvate dehydrogenase B. Biomed. Pharmacother. 2017; 90: 278-286. 10.1016/j.biopha.2017.03.068 Peng Y., Croce C.M. The role of MicroRNAs in human cancer. Signal Transduct. Target. Ther. 2016; 1: 15004. 10.1038/sigtrans.2015.4. 29263891 Macfarlane L.A., Murphy P.R. MicroRNA: Biogenesis, Function and Role in Cancer. Curr. Genom. 2010; 11: 537-561. 10.2174/138920210793175895. 21532838 Bartel D.P. MicroRNAs: Genomics, biogenesis, mechanism, and function. Cell. 2004; 116: 281-297. 10.1016/S0092-8674(04)00045-5 Subramaniam S., Jeet V., Clements J.A., Gunter J.H., Batra J. Emergence of MicroRNAs as Key Players in Cancer Cell Metabolism. Clin. Chem. 2019. 10.1373/clinchem.2018.299651. 31101638 Dong J., Xiao D., Zhao Z., Ren P., Li C., Hu Y., Shi J., Su H., Wang L., Liu H. Epigenetic silencing of microRNA-137 enhances ASCT2 expression and tumor glutamine metabolism. Oncogenesis. 2017; 6: e356. 10.1038/oncsis.2017.59 Anderson N.M., Mucka P., Kern J.G., Feng H. The emerging role and targetability of the TCA cycle in cancer metabolism. Protein Cell. 2018; 9: 216-237. 10.1007/s13238-017-0451-1 Choi Y.K., Park K.G. Targeting Glutamine Metabolism for Cancer Treatment. Biomol. Ther. 2018; 26: 19-28. 10.4062/biomolther.2017.178 Tanaka H., Sasayama T., Tanaka K., Nakamizo S., Nishihara M., Mizukawa K., Kohta M., Koyama J., Miyake S., Taniguchi M. MicroRNA-183 upregulates HIF-1alpha by targeting isocitrate dehydrogenase 2 (IDH2) in glioma cells. J. Neurooncol. 2013; 111: 273-283. 10.1007/s11060-012-1027-9 Chu B., Wu T., Miao L., Mei Y., Wu M. MiR-181a regulates lipid metabolism via IDH1. Sci. Rep. 2015; 5: 8801. 10.1038/srep08801. 25739786 Galluzzi L., Morselli E., Vitale I., Kepp O., Senovilla L., Criollo A., Servant N., Paccard C., Hupe P., Robert T. miR-181a and miR-630 regulate cisplatin-induced cancer cell death. Cancer Res. 2010; 70: 1793-1803. 10.1158/0008-5472.CAN-09-3112. 20145152 Li Q.J., Chau J., Ebert P.J., Sylvester G., Min H., Liu G., Braich R., Manoharan M., Soutschek J., Skare P. miR-181a is an intrinsic modulator of T cell sensitivity and selection. Cell. 2007; 129: 147-161. 10.1016/j.cell.2007.03.008. 17382377 Bader A.G., Brown D., Winkler M. The promise of microRNA replacement therapy. Cancer Res. 2010; 70: 7027-7030. 10.1158/0008-5472.CAN-10-2010 Hsu P.D., Lander E.S., Zhang F. Development and applications of CRISPR-Cas9 for genome engineering. Cell. 2014; 157: 1262-1278. 10.1016/j.cell.2014.05.010 Doudna J.A., Charpentier E. Genome editing. The new frontier of genome engineering with CRISPR-Cas9. Science. 2014; 346: 1258096. 10.1126/science.1258096 Gebler C., Lohoff T., Paszkowski-Rogacz M., Mircetic J., Chakraborty D., Camgoz A., Hamann M.V., Theis M., Thiede C., Buchholz F. Inactivation of Cancer Mutations Utilizing CRISPR/Cas9. J. Natl. Cancer Inst. 2017; 109. 10.1093/jnci/djw183 Flavahan W.A., Drier Y., Liau B.B., Gillespie S.M., Venteicher A.S., Stemmer-Rachamimov A.O., Suva M.L., Bernstein B.E. Insulator dysfunction and oncogene activation in IDH mutant gliomas. Nature. 2016; 529: 110-114. 10.1038/nature16490 Fogleman S., Santana C., Bishop C., Miller A., Capco D.G. CRISPR/Cas9 and mitochondrial gene replacement therapy: Promising techniques and ethical considerations. Am. J. Stem Cells. 2016; 5: 39-52 Haapaniemi E., Botla S., Persson J., Schmierer B., Taipale J. CRISPR-Cas9 genome editing induces a p53-mediated DNA damage response. Nat. Med. 2018; 24: 927-930. 10.1038/s41591-018-0049-z

By Jubert Marquez; Jessa Flores; Amy Hyein Kim; Bayalagmaa Nyamaa; Anh Thi Tuyet Nguyen; Nammi Park and Jin Han

Reported by Author; Author; Author; Author; Author; Author; Author

Titel:
Rescue of TCA Cycle Dysfunction for Cancer Therapy
Autor/in / Beteiligte Person: Marquez, Jubert ; Ah, Kim ; Park, N ; Nyamaa, Bayalagmaa ; Han, Jin ; Att, Nguyen ; Flores, J
Link:
Zeitschrift: Journal of Clinical Medicine, Jg. 8 (2019-12-06), S. 2161-2161
Veröffentlichung: MDPI AG, 2019
Medientyp: unknown
ISSN: 2077-0383 (print)
DOI: 10.3390/jcm8122161
Schlagwort:
  • IDH
  • 0301 basic medicine
  • Cell
  • Review
  • SDH
  • Oxidative phosphorylation
  • Mitochondrion
  • MDH
  • FH
  • 03 medical and health sciences
  • 0302 clinical medicine
  • microRNA
  • medicine
  • cancer
  • CRISPR/Cas9
  • miRNA
  • chemistry.chemical_classification
  • business.industry
  • Cancer
  • General Medicine
  • medicine.disease
  • TCA
  • Cell biology
  • mitochondria
  • Citric acid cycle
  • 030104 developmental biology
  • medicine.anatomical_structure
  • Enzyme
  • chemistry
  • 030220 oncology & carcinogenesis
  • business
  • Function (biology)
Sonstiges:
  • Nachgewiesen in: OpenAIRE
  • Rights: OPEN

Klicken Sie ein Format an und speichern Sie dann die Daten oder geben Sie eine Empfänger-Adresse ein und lassen Sie sich per Email zusenden.

oder
oder

Wählen Sie das für Sie passende Zitationsformat und kopieren Sie es dann in die Zwischenablage, lassen es sich per Mail zusenden oder speichern es als PDF-Datei.

oder
oder

Bitte prüfen Sie, ob die Zitation formal korrekt ist, bevor Sie sie in einer Arbeit verwenden. Benutzen Sie gegebenenfalls den "Exportieren"-Dialog, wenn Sie ein Literaturverwaltungsprogramm verwenden und die Zitat-Angaben selbst formatieren wollen.

xs 0 - 576
sm 576 - 768
md 768 - 992
lg 992 - 1200
xl 1200 - 1366
xxl 1366 -