Skip to main content
Erschienen in: Malaria Journal 1/2017

Open Access 01.12.2017 | Research

Suppression of experimental cerebral malaria by disruption of malate:quinone oxidoreductase

verfasst von: Mamoru Niikura, Keisuke Komatsuya, Shin-Ichi Inoue, Risa Matsuda, Hiroko Asahi, Daniel Ken Inaoka, Kiyoshi Kita, Fumie Kobayashi

Erschienen in: Malaria Journal | Ausgabe 1/2017

Abstract

Background

Aspartate, which is converted from oxaloacetate (OAA) by aspartate aminotransferase, is considered an important precursor for purine salvage and pyrimidine de novo biosynthesis, and is thus indispensable for the growth of Plasmodium parasites at the asexual blood stages. OAA can be produced in malaria parasites via two routes: (i) from phosphoenolpyruvate (PEP) by phosphoenolpyruvate carboxylase (PEPC) in the cytosol, or (ii) from fumarate by consecutive reactions catalyzed by fumarate hydratase (FH) and malate:quinone oxidoreductase (MQO) in the mitochondria of malaria parasites. Although PEPC-deficient Plasmodium falciparum and Plasmodium berghei (rodent malaria) parasites show a growth defect, the mutant P. berghei can still cause experimental cerebral malaria (ECM) with similar dynamics to wild-type parasites. In contrast, the importance of FH and MQO for parasite viability, growth and virulence is not fully understood because no FH- and MQO-deficient P. falciparum has been established. In this study, the role of FH and MQO in the pathogenicity of asexual-blood-stage Plasmodium parasites causing cerebral malaria was examined.

Results

First, FH- and MQO-deficient parasites were generated by inserting a luciferase-expressing cassette into the fh and mqo loci in the genome of P. berghei ANKA strain. Second, the viability of FH-deficient and MQO-deficient parasites that express luciferase was determined by measuring luciferase activity, and the effect of FH or MQO deficiency on the development of ECM was examined. While the viability of FH-deficient P. berghei was comparable to that of control parasites, MQO-deficient parasites exhibited considerably reduced viability. FH activity derived from erythrocytes was also detected. This result and the absence of phenotype in FH-deficient P. berghei parasites suggest that fumarate can be metabolized to malate by host or parasite FH in P. berghei-infected erythrocytes. Furthermore, although the growth of FH- and MQO-deficient parasites was impaired, the development of ECM was suppressed only in mice infected with MQO-deficient parasites.

Conclusions

These findings suggest that MQO-mediated mitochondrial functions are required for development of ECM of asexual-blood-stage Plasmodium parasites.
Begleitmaterial
Additional file 1. Fumarate cycle in Plasmodium falciparum. Fumarate, which is generated via purine biosynthesis, is converted into malate by fumarate hydratase (FH) [11]. Then, malate is converted to oxaloacetate (OOA) by malate:quinone oxidoreductase (MQO) [12]; the oxidation of malate to OOA generates ubiquinol (UQH2), which feeds the electron transport chain [11, 12]. Two of the eight mitochondrial TCA cycle enzymes, FH and MQO, may be essential for survival of asexual-blood-stage P. falciparum [9]. Note: MQO is conserved among all apicomplexan parasites, including all Cryptosporidium species [29].
Additional file 3. Generation of FH- and MQO-deficient Plasmodium berghei. SK-1 vector (A) and Sk-1-luc2 vector (B). Restriction sites of NheI and BglII restriction enzymes were shown. Schematic representation of gene-targeting vectors (A and B). Luciferase (luc2)-expressing cassette was introduced into target gene by double-crossover homologous recombination. Arrows (F1, F2, M1 and M2) denote primers specific for the 5′ and 3′ regions of the target gene (see Additional file 2). (A) Introduction of luc2-expressing cassette into the fh locus (PBANKA_082810) of P. berghei parasites. Proper integration was confirmed using primers specific for fh (WT, 3.0 kbp; Δfh, 6.8 kbp) for three cloned transfected parasites. (B) Introduction of luc2-expressing cassette into the mqo locus (PBANKA_111630) of P. berghei parasites. Proper integration was confirmed using primers specific for mqo (WT, 3.0 kbp; Δmqo, 7.0 kbp) for three cloned transfected parasites.
Additional file 4. Deficiency of FH and MQO has no effect on gametocyte production. Blood was obtained from infected mice showing 3% parasitaemia and cultured for 22 h under standardized in vitro culture conditions. Then, mature gametocytes and schizonts were collected by Nycodenz density-gradient centrifugation. (A and B) Expression of gametocyte-specific genes. mdv-1/peg3 [21] and g377 [22] were subjected to semi-quantitative RT–PCR using specific primers (see Additional files 2, 3). The hsp70 was used as a positive control. Samples treated with DNase-treated RNA template (hsp70 (-)) were used as a negative control that is the control of eventual DNA contamination of the RNA preparations. Experiments were performed in duplicate and representative data are shown. (C) Control and Δfh parasites-infected erythrocytes cultured for 22 h. (D) Control and Δmqo parasites-infected erythrocytes cultured for 22 h. White arrows indicate representative mature gametocytes. The scale bars indicate 20 μm. Note that sex-specific features such as nuclear enlargement, the distribution of pigment granules throughout the cytoplasm and enlargement of the cells are observed in both Δfh- and Δmqo-parasite cultures just the same as reported by Mons [23].
Additional file 5. Parasitaemia is suppressed by deficiency of parasite MQO but not FH. Bar graph of parasitaemia in Fig. 5A. Parasitaemia on days 3–7 post-infection were shown. Asterisks indicate a statistically significant difference (*, vs. control).
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1186/​s12936-017-1898-5) contains supplementary material, which is available to authorized users.
Abkürzungen
DCIP
2,6-dichlorophenolindophenol
APAD+
3-acetylpyridine-adenine dinucleotide, oxidized form
ATP
adenosine triphosphate
AAT
aspartate aminotransferase
DTC
dicarboxylate–tricarboxylate carrier homolog
ETC
electron transport chain
ECM
experimental cerebral malaria
FH
fumarate hydratase
hdhfr
human dihydrofolate reductase-thymidylate synthase
MQO
malate:quinone oxidoreductase
NBT
nitroblue tetrazolium
OAA
oxaloacetate
PEP
phosphoenolpyruvate
PEPC
phosphoenolpyruvate carboxylase
TCA
tricarboxylic acid
UQH2
ubiquinol
UQ
ubiquinone
mATPβ
β subunit gene of mitochondrial ATP synthase

Background

Plasmodium species are among the most important mosquito-borne pathogens worldwide, and cause an estimated 212 million malaria cases and 429,000 deaths due to malaria per year [1]. When an infected mosquito takes a blood meal, a small number of Plasmodium sporozoites are injected into the host’s bloodstream. The injected sporozoites invade hepatocytes and produce merozoites. These merozoites are released into the bloodstream and invade erythrocytes, in which the vast majority multiply asexually; only a small subset of parasites differentiate into sexual precursor cells (i.e., male and female gametocytes [2]).
The metabolic pathways in Plasmodium parasites differ from those of their host. These parasites use nutrients obtained from the host [3] and, to sustain parasite growth, adenosine triphosphate (ATP) is produced (mainly by glycolysis). Although Plasmodium spp. possesses all of the genes necessary for the tricarboxylic acid (TCA) cycle, [4] and most of the genes needed for electron transport chain (ETC) enzymes, asexual-blood-stage malaria parasites rely mainly on cytosolic glycolysis with limited contribution from mitochondrial oxidative phosphorylation for ATP synthesis [5, 6]. Several reports [79] have demonstrated that the TCA cycle is not essential for survival of asexual-blood-stage parasites, but is required for survival of mosquito-stage parasites. However, two of the eight mitochondrial TCA cycle enzymes, fumarate hydratase (FH) and malate:quinone oxidoreductase (MQO), could not be genetically ablated in asexual-blood-stage Plasmodium falciparum, suggesting that these two enzymes are promising drug targets [9].
In this regard, the “fumarate cycle” should be noted (see Additional file 1). The purine salvage pathway is an important source of fumarate, which is generated as a by-product of the adenylosuccinate lyase reaction in Plasmodium [10] (see Additional file 1). This fumarate can then be converted into malate by the malarial FH [11], and then to OAA by MQO in mitochondria [12] (see Additional file 1). OAA generated by MQO is converted to aspartate by aspartate aminotransferase (AAT) in the cytosol, which feeds the purine salvage pathway, through which fumarate is regenerated, in a process termed the fumarate cycle [11] (see Additional file 1). The oxidation of malate to oxaloacetate by MQO is coupled to reduction of ubiquinone (UQ) to form ubiquinol (UQH2), which then feeds into the ETC at complex III [11, 12]. This purine salvage pathway, TCA cycle and ETC network suggests the presence of intense metabolic cross-talk in Plasmodium parasites [11].
OAA can be produced in malaria parasites from (i) fumarate by consecutive reactions catalyzed by FH and MQO in the mitochondria of malaria parasites, as described above, or from (ii) phosphoenolpyruvate (PEP; common in plants and bacteria) by phosphoenolpyruvate carboxylase (PEPC) in the cytosol of the parasite. PEPC-deficient P. falciparum has a severe growth defect. In contrast, growth of PEPC-deficient P. falciparum is partially rescued by supplementation of cultures with a high concentration of fumarate or malate [13]. Interestingly, PEPC-deficient Plasmodium berghei cause severe cerebral malaria, with dynamics similar to those caused by wild-type parasites [14]. However, the importance of FH and MQO for asexual-stage parasite viability and growth in cerebral malaria is unclear.
In this study, FH and MQO in P. berghei (strain ANKA), which is the aetiologic agent of experimental cerebral malaria (ECM) in rodents were focused. To investigate the roles of FH and MQO in the viability and growth of malaria parasites, a luciferase-expressing cassette was introduced into the mqo and fh loci in the genome of P. berghei, and the viability of fh-disrupted (Δfh) and mqo-disrupted (Δmqo) parasites was evaluated by measuring luciferase activities. Moreover, the effect of FH and MQO deficiency on the development of ECM caused by P. berghei were assessed.

Methods

Mice

Female C57BL/6J mice 5- to 6-weeks old were purchased from CLEA Japan INC (Tokyo, Japan). The experiments were approved by the Experimental Animal Ethics Committee of Kyorin University School of Medicine, Tokyo, and all experimental animals were kept at the animal facility in a specific-pathogen-free unit with sterile bedding, food and water.

DNA constructs

The SK-1 construct contained a selection cassette consisting of the green fluorescent protein gene (gfp) and a pyrimethamine resistance gene, human dihydrofolate reductase-thymidylate synthase (hdhfr) [15]. The expression of gfp and hdhfr is controlled by hsp70 (PBANKA_071190) and elongation factor-1 (PBANKA_113340) promoters, respectively. Plasmid (pLG4.10[luc2]) containing luciferase gene (luc2) was purchased from Promega (Madison, WI, USA). To generate luciferase-expressing cassette, luc2 was amplified by PCR using specific primers (see Additional files 2, 3). The PCR product of luc2 and SK-1 construct were cleaved using the NheI and BglII restriction enzymes, and the gfp of SK-1 was replaced with luc2. The gene-targeting vectors were prepared by PCR [16]. Briefly, the 5′ and 3′ flanking regions the ORF of the target genes, p230 [17], fh locus (PBANKA_082810) and mqo locus (PBANKA_111630), were amplified by PCR. The PCR products were annealed to either side of luciferase expressing cassette and amplified by PCR using gene-specific primers (see Additional file 2).

Parasites and infections

Plasmodium berghei ANKA strain is a high-virulence strain and the parasites, which had been cloned by limiting dilution, were obtained from Dr. W. P. Weidanz (University of Wisconsin–Madison, Madison, WI, USA). Erythrocytes infected with P. berghei parasites were transferred to RPMI1640 medium supplemented with 25% (v/v) FBS, 0.05 mg/mL Penicillin, 0.05 mg/mL Streptomycin and then incubated for 18 or 22 h under condition of 90% N2, 5% CO2 and 5% O2. Mature schizonts and gametocytes were collected by Nycodenz density-gradient centrifugation [18]. Transformations were performed using the Amaxa Basic Parasite Nucleofector Kit (Amaxa GmbH, Cologne, Germany) according to the manufacturer’s protocol and luciferase-expressing cassette was introduced into the ORF of targeted gene by double-crossover homologous recombination [18]. Briefly, 5 × 106 to 5 × 107 purified P. berghei mature schizonts were mixed with 100 μL of Nucleofector solution containing 5 μg of a gene-targeting vector. Transfections were then completed using the Amaxa Nucleofector electroporation program U-33. Transfected parasites were then injected intravenously (i.v.) into naïve C57BL/6 recipient mice. At 30 h post-injection, transfected parasites were selected by addition of pyrimethamine to the drinking water of infected mice. After parasitaemia returned to detectable levels post-selection, transfected parasites were cloned by limiting dilution, after which a single parasite was injected into a mouse to ensure a clonally pure population. Cloned transfected parasites were stored as frozen stocks in liquid nitrogen. Erythrocytes infected with transfected parasites were generated in donor mice inoculated intraperitoneally with each frozen stock of parasites. The donor mice were monitored for parasitaemia daily and bled for experimental infection in ascending periods of parasitaemia. Experimental mice were infected intravenously with 1 × 104 infected erythrocytes or 5 × 106 to 5 × 107 purified mature schizonts of a given parasite strain.

Parasitaemia

Blood was observed by microscopic examination of methanol-fixed tail blood smears stained with 3% (w/v) Giemsa diluted with Sörensen’s phosphate buffer, pH 7.2, for 45 min. The number of infected erythrocytes in 250 erythrocytes was enumerated when parasitaemia exceeded 10%, whereas 1 × 104 erythrocytes were examined when mice showed lower parasitaemia. The percentage of parasitaemia was calculated as follows: [(Number of infected erythrocytes)/(Total number of erythrocytes)] × 100.

Genomic PCR

To generate luciferase-expressing cassette and confirm the introduction of Luc2-expressing cassette into target gene, genomic PCR was performed using a 25 μL PCR mixture containing 1×TaKaRa PrimeSTAR GXL buffer (TaKaRa, Shiga, Japan), 2.5 mM dNTPs, 0.5 μL of DNA, 5 U/μL TaKaRa PrimeSTAR GXL DNA polymerase (TaKaRa), and PCR primers (0.25 μM). Thirty-five cycles of PCR were performed on a C1000 thermal cycler (Bio-Rad, Hercules, CA, USA). Each cycle consisted of denaturation at 98 °C for 15 s, annealing at 55 °C for 15 s, and extension at 68 °C for 1–6 min. PCR products were then analysed on a 1% (w/v) agarose gel, and stained with ethidium bromide.

Semi-quantitative RT-PCR

Blood was obtained from infected mice exhibiting 2–5% parasitaemia. Total RNA was isolated from blood containing 5 × 106 or 1 × 107 infected erythrocytes using Isogen (Nippon Gene, Tokyo, Japan) according to the manufacturer’s protocol. Total RNA was then treated with DNase and reverse-transcribed using murine leukaemia virus reverse transcriptase (Applied Biosystems, Carlsbad, CA, USA) with random hexamer primers under the following conditions: 70 °C for 10 min, 25 °C for 10 min, and 42 °C for 30 min. The reaction was terminated by heating at 99 °C for 5 min, and the resulting cDNA products were stored at −20 °C until required. All PCR reactions were run in a 25 μL volume consisting of 1×TaKaRa Ex Taq buffer, 2.5 mM dNTP, 1 μL of cDNA products, 5 U/μL TaKaRa Ex Taq DNA polymerase, and PCR primers (0.25 μM); a list of primers used for semi-quantitative RT-PCR can be found in Additional file 2. PCR reactions were performed on a C1000 thermal cycler (Bio Rad) for 30 cycles under the following conditions: denaturation at 95 °C for 30 s, annealing at 55 °C for 30 s and extension at 72 °C for 45 s. Products were resolved on a 2% (w/v) agarose gel, and stained with ethidium bromide. Samples with DNase-treated RNA template were used as the negative control.

Preparation of mitochondria

To remove leukocytes, the blood was mixed with same volume of PBS and passed over a Plasmodipur Filter (EuroProxima, Arnhem, Netherlands). Erythrocytes were washed twice with RPMI1640 medium by centrifugation at 4 °C at 800×g, for 5 min and then transferred to RPMI1640 medium supplemented with 25% (v/v) FBS, 0.05 mg/mL Penicillin, 0.05 mg/mL Streptomycin. Then erythrocytes were incubated for 16 h under condition of 90% N2, 5% CO2 and 5% O2. Stages of parasite were mainly schizonts as confirmed by Giemsa staining. Infected parasites were collected by centrifugation at 4 °C at 800×g, for 5 min. Crude mitochondria of P. berghei were prepared as described previously [19].

Measurement of MQO activity

MQO activity assay was performed at 25 °C with V-660 spectrophotometer (JASCO) and measured in 1 mL of the reaction mixture containing 20 µg of mitochondrial fraction, 45 µM 2,6-dichlorophenolindophenol (DCIP), 100 µM ubiquinone-2 and 2 mM KCN in 50 mM potassium phosphate buffer, pH 8.0 by recording the decrease in absorbance due to DCIP reduction at 600 nm (ε600 = 21/mM/cm) after the reaction was initiated by adding 10 mM sodium malate.

Measurement of FH activity

FH activity was performed 37 °C with Benchmark Plus microplate spectrophotometer (Bio-Rad) and measured in 200 µL of the reaction mixture containing 20 µg of mitochondrial fraction, 0.25% (v/v) Triton X-100, 50 µg/mL 3-acetylpyridine-adenine dinucleotide, oxidized form (APAD+) (Oriental Yeast, Japan), 1 U/mL diaphorase, 0.2 mg/mL nitroblue tetrazolium (NBT) and 2 U/mL malate dehydrogenase from pig heart (Oriental Yeast) in 100 mM Tris–HCl buffer pH 8.0 by recording the absorbance change of NBT at 530 nm (ε530 = 3.6 × 104/mM/cm) after the reaction was initiated by adding 10 mM sodium fumarate.

Luciferase assay for evaluation of the viability of malaria parasites

To monitor the viability of fh- and mqo-disrupted P. berghei by luciferase activities, luciferase-expressing cassette was introduced into fh locus (PBANKA_082810) and mqo locus (PBANKA_111630) in the genome of P. berghei (see Additional file 3). Moreover, luciferase-expressing cassette was also introduced into p230 locus (PBANKA_030600), which is not an essential gene in the complete life cycle of P. berghei [17], and the resultant was used as the control parasite in the following experiments. Blood was obtained from C57BL/6 mice exhibiting 2–5% parasitaemia, and diluted with RPMI 1640 medium to a final concentration of 1 × 107 infected erythrocytes/well in a 96-well plate. Then, infected erythrocytes were cultured for 3 h in RPMI 1640 medium supplemented with 5 mM sodium fumarate (Sigma-Aldrich Co, Missouri, USA), 5 mM sodium malate (Sigma-Aldrich Co) or control PBS. After cultivation, d-luciferin (1.5 mM; Promega) was added to each well, and luminescence measured using an SH9000 luminometer (Hitachi High-Technologies Corporation, Tokyo, Japan) at 5 min after addition of d-luciferin. Results are presented as relative fluorescence unit (RFU) fold change compared with control PBS at 5 min after addition of d-luciferin. Also, blood was obtained from C57BL/6 mice on 16 h after injection of schizonts, and diluted with RPMI 1640 medium to a final concentration of 5 × 106 infected erythrocytes/well in a 96-well plate. After cultivation, d-luciferin (1.5 mM; Promega) was added to each well, and luminescence measured using an SH9000 luminometer at 20 min after addition of d-luciferin. Results are presented as absolute luciferase activity in counts per second (cps).

Determination of mitochondrial membrane potential

To investigate whether mitochondrial membrane potential (MMP) in malaria parasites is compromised by deficiency of FH or MQO, MMP in trophozoites of control, Δfh and Δmqo parasites was determined using MMP-sensitive fluorochrome Mitotracker® Red CMXRos (Invitrogen, Darmstadt, Germany). Inner membrane and nuclear DNA were then stained with the BODIPY FL C16 (Invitrogen) and Hoechst 33342 (Invitrogen), respectively. Mitotracker® Red CMXRos was added to medium with a concentration of 100 nM and incubated for 30 min at 37 °C. Then, BODIPY FL C16 and Hoechst 33342 were added to medium with a concentration of 100 nM and of 1 µg/mL for 10 min at 37 °C. For all staining of malaria parasites, the staining medium was removed after the incubation period and the fresh medium was added. The Bright field and fluorescence microscopy images were photographed at 1000× magnification using an All-in-One Fluorescence Microscope (BZ9000; KEYENCE Japan, Osaka, Japan). Mean fluorescence intensity (MFI) of MitoTracker in the photographs was analysed using BZ-II Analyzer software (KEYENCE Japan).

Examination of the blood–brain barrier

It is known that breakdown of the blood–brain barrier is an indicator of ECM. When P. berghei-infected mice develop ECM, Evans blue is injected i.v. and the brain is stained as a result of extravasation of the dye [20]. Mice were injected i.v. with 0.2 mL of 1% (w/v) Evans blue (Wako, Osaka, Japan) on days 7 and 14 post-infection. Mice were euthanized and brains perfused with PBS 1 h later. After the brains were removed and photographed they were weighed and placed in 2 mL formamide (Wako, Osaka, Japan) at 37 °C for 48 h to extract the Evans blue dye. Absorbance was measured at λ = 620 nm with a Multiscan FC microplate reader (Thermo Fisher Scientific Inc., Waltham, USA). The Evans blue concentration was calculated from a standard curve and is expressed as µg of Evans blue per g of brain.

Statistical analysis

Student’s t test was performed using Statcel (OMS Ltd., Saitama, Japan). Statistically significant differences were defined as a value of p < 0.05.

Results

Generation of FH-disrupted and MQO-disrupted Plasmodium berghei

To investigate the role of FH and MQO in malaria parasite growth in erythrocytes, fh- and mqo-disrupted P. berghei were generated. The luciferase mRNA level in fh-disrupted (Δfh) and mqo-disrupted (Δmqo) parasites was comparable to that in control parasites (Fig. 1a, b). Moreover, fh and mqo mRNAs were not detected in Δfh and Δmqo parasites, respectively (Fig. 1a, b). These findings demonstrate that the luciferase-expressing cassette was successfully introduced into fh and mqo locus.
In Δfh parasite-infected erythrocytes, the specific activity of FH was 0.99 ± 0.12 nmol/min/mg, which was 58.5 and 60.0% lower compared to control-infected and Δmqo parasite-infected erythrocytes (1.69 ± 0.91 and 1.65 ± 0.54 nmol/min/mg, respectively) (Fig. 1c). As FH activity was detected in uninfected erythrocytes, the FH activity in Δfh parasite-infected erythrocytes seems to be derived from host FH (Fig. 1c). The MQO activity in Δmqo parasite-infected erythrocytes was completely abrogated, although high activities in control parasite- and Δfh parasite-infected erythrocytes were detected (Fig. 1d).

Inhibition of parasite growth during late trophozoites and schizont stage by deficiency of FH and MQO

The effect of deficiency of FH and MQO on parasite growth in vivo were investigated (Fig. 2a). Δfh and Δmqo parasites showed a growth pattern similar to control parasites, until 18 h post-inoculation of schizonts (Fig. 2a). However, at 24 h post-inoculation, the proportions of late trophozoites (stage 3) of Δfh and Δmqo parasites were considerably higher than in control parasites (Fig. 2a). These findings suggest that parasite growth during late trophozoites and schizont stage is delayed by deficiency of FH or MQO.
To evaluate the production of gametocytes in Δfh and Δmqo parasites, the mRNA levels of the gametocyte-specific proteins MDV1/PEG3 and g377 were examined. mdv-1/peg3 and g377 mRNAs can be detected from the early phase of gametocytogenesis [21, 22]. As results, gametocyte-specific gene expression during asexual-blood-stage Δfh and Δmqo parasites was comparable to that of control parasites (see Additional file 4A, B). Nuclear enlargement, the distribution of pigment granules throughout the cytoplasm and enlargement of cells were observed in a 22 h culture of Δfh and Δmqo parasites (see Additional file 4C, D). These three features are gametocyte-specific characteristics [23]. Taken together, these results suggest that the development of female and male gametocytes of malaria parasites was not affected by deficiency of FH and MQO.

Effect of FH and MQO deficiency on the viability of malaria parasites

ATP plays a central role in energy transduction in both eukaryotic and prokaryotic cells, and is produced in all metabolically active cells. The luciferase–luciferin system, in which luciferase reacts with d-luciferin in the presence of ATP, an oxygen molecule, and magnesium ions to produce luminescence, has facilitated investigation of parasite viability [24]. To examine the effect of deficiency of FH and MQO on parasite viability, schizonts of control, Δfh and Δmqo parasites were inoculated into mice. Trophozoites were obtained at 8 and 16 h post-inoculation and their luciferase activities were measured. The absolute luciferase activity values of Δfh parasites were comparable to those of control parasites (Fig. 2b). In contrast, the luciferase activities of Δmqo parasites were significantly reduced, by ~50% compared with those of control and Δfh parasites (Fig. 2b). These findings suggest that the viability of asexual-blood-stage malaria parasites is decreased by deficiency of MQO but not FH.
To investigate the effect of FH and MQO deficiency on mitochondrial function, the study compared the mitochondrial membrane potential (MMP) of control, Δfh and Δmqo parasites by monitoring MitoTracker uptake. MMP was not affected by deficiency of FH and MQO (Fig. 3), suggesting that the reduction of parasite viability caused by MQO deficiency is independent of defects in mitochondrial integrity.

Effect of fumarate and malate on the viability of control, Δfh and Δmqo parasites

Results that luciferase activities were not affected by deficiency of FH suggest cooperation of the host- and parasite-derived FH in fumarate metabolism. Therefore, the effect of addition of fumarate or malate to the culture on the viability of control, Δfh and Δmqo parasites were investigated using the luciferase–luciferin system. In the culture of control parasites, luciferase activities were increased by addition of fumarate or malate compared with the control (Fig. 4). The luciferase activities of Δfh parasites and Δmqo parasites were also increased in culture medium supplemented with fumarate or malate (Fig. 4). Therefore, enriched metabolic substitutions by host-derived enzymes, such as FH, in erythrocytes may be associated with increased viability of Δfh and Δmqo parasites in culture medium supplemented with fumarate or malate.

Suppression of ECM development by deficiency of MQO but not FH

Plasmodium berghei strain ANKA is the causative agent of ECM, in which neurologic signs and histopathological findings, such as haemorrhages and sequestration of infected erythrocytes within cerebral microvessels, are observed [20]. However, the relationship between parasite growth and pathogenicity in ECM is unclear. Therefore, the effect of FH or MQO deficiency on ECM were examined. Mice infected with Δfh and Δmqo parasites showed lower levels of parasitaemia than mice infected with control parasites until days 5 and 7 post-infection, respectively (Fig. 5a; see Additional file 5). All mice infected with Δfh parasites showed neurologic signs and died on days 6–7 post-infection (Fig. 5b); this is similar to the effect in control mice. In contrast, the survival of mice infected with Δmqo parasites was prolonged compared to that of mice infected with control parasites or Δfh parasites (Fig. 5b). The mice infected with Δmqo parasites ultimately died via anaemia within days 25–30 post-infection.
Breakdown of the blood–brain barrier is an indicator of ECM [20]; therefore, breakdown of the blood–brain barrier in mice infected with control, Δfh or Δmqo parasites were investigated by assessing extravasation of Evans blue in the brain. The levels of extravasation of Evans blue in the brains of Δfh-infected mice on day 7 post-infection were comparable to those in mice infected with control parasites (Fig. 5c). In contrast, the levels of extravasation of Evans blue in the brains of Δmqo-infected mice on days 7 and 14 post-infection were markedly lower than those in mice infected with control or Δfh parasites, and were comparable to those of uninfected mice (Fig. 5c). These results suggest that the ECM caused by malaria parasites is suppressed by deficiency of parasite MQO, but not FH.

Discussion

This study demonstrated that the growth of P. berghei was delayed by deficiency of FH and MQO, and that the development of ECM was suppressed in mice infected with Δmqo parasites. This suggests an essential role for MQO in the fumarate cycle. OAA produced by MQO in mitochondria is transported to the cytosol by a dicarboxylate–tricarboxylate carrier homolog (DTC) [25] and converted to aspartate, which is necessary for the purine salvage pathway (see Additional file 1) [11].
The viability and growth of P. berghei was reduced by deficiency of MQO. MQO reduces UQ to UQH2, which is associated with chemiosmotic gradient maintenance and ATP synthesis via ETC and oxidative phosphorylation in the mitochondria [11, 12]. Previous studies have suggested that the gene encoding the β subunit of mitochondrial ATP synthase (mATPβ) could be disrupted in P. berghei [26]. The growth of asexual P. berghei parasites was delayed by deficiency of mATPβ; however, it is unknown whether mice infected with mATPβ-deficient P. berghei develop ECM [26]. These findings suggest that MQO-mediated mitochondrial function is required for full virulence of asexual-blood-stage Plasmodium parasites.
In addition, the viability of Δfh parasites was comparable to that of control parasites. Detection of host-derived FH activity suggests that host-derived FH functions in erythrocytes infected with Δfh parasites are involved in maintain fumarate metabolism. FH is localized in the mitochondria and cytosol in all eukaryotes [27, 28]. In Δfh parasites, fumarate, which is generated via the purine nucleotide cycle, may be secreted into reticulocytes and converted to malate by host-derived FH; malate ultimately enters the fumarate cycle via MQO in malaria parasites.
In this study, fumarate or malate supplementation increased the luciferase activity of FH- and MQO-deficient parasites, as well as of control parasites. These results suggest that the increase in viability is independent of malarial FH and MQO. In uninfected erythrocytes, FH and cytosolic malate dehydrogenase are present, which represents an alternative OAA biosynthesis pathway. This increased aspartate production, via AAT and pyrimidine de novo biosynthesis, enhances the growth of Plasmodium parasites. This hypothesis is supported by the fact that establishment of AAT-deficient malaria parasites has been unsuccessful [14].
This study suggests that FH and MQO are not essential for survival of rodent malaria parasites that mainly infects reticulocytes; indeed, cultures of FH- and MQO-deficient P. falciparum could not be established [9]. Srivastava et al. [14] reported that the metabolic profile in reticulocytes is enriched compared with that in mature human and mouse erythrocytes. Therefore, the interaction between host and parasite metabolism is enhanced in reticulocytes compared to in mature erythrocytes.
MQO is conserved among all apicomplexan parasites, including Cryptosporidium and Perkinsus, which are early branching groups of chromalveolates (apicomplexa and dinoflagellates, respectively). Despite the absence of genes encoding TCA cycle enzymes in the genome of Cryptosporidium parvum, few mitochondrial enzymes—such as MQO, type-II NADH dehydrogenase and cyanide-insensitive alternative oxidase—are conserved in the C. parvum mitosome [29]. This observation led to the hypothesis that MQO plays a structural rather than metabolic role in these parasites [9]. The inner membrane structure of mouse mitochondria is regulated by MMP [30]. The effect of MQO on MMP and inner membrane structure was evaluated using MitoTracker® Red CMXRos and BODIPY FL C16, respectively. However, no obvious change in MMP or inner membrane structure was detected in Δmqo parasites, indicating that the primary role of MQO is metabolic rather than structural, at least in P. berghei parasites.

Conclusion

Anti-malarial atovaquone kills both the blood and liver stages of Plasmodium parasites [31], and atovaquone-resistant parasites cannot be transmitted to other hosts [32]; this indicates that mitochondrial function is necessary for the viability and growth of Plasmodium parasites at all lifecycle stages. In P. falciparum, a functional ETC is important for survival of blood-stage parasites [31], while both the TCA cycle and the ETC are critical for parasite development at the insect stage [7]. In Plasmodium parasites, the enzymes for pyrimidine salvage and purine de novo pathways are not conserved; therefore, these parasites rely solely on the pyrimidine de novo and purine salvage pathways for pyrimidine and purine production. MQO is a key enzyme in supplying OAA for the fumarate cycle, which produces aspartate in the cytosol to feed the pyrimidine de novo and purine salvage pathways [9, 13]. Hence, MQO is a trifunctional enzyme involved in the TCA cycle, ETC and fumarate cycle. The finding that MQO, but not FH, is involved in both the viability and growth of asexual-blood-stage parasites and development of ECM suggest it to be a potential target for the treatment and prevention of the transmission of severe malaria.

Authors’ contributions

MN, KK and FK designed research; MN, KK and RM performed research; MN, KK, S-II, HA, DKI, KK, and FK analysed data; and MN, KK, DKI, KK and FK wrote the paper. All authors read and approved the final manuscript.

Acknowledgements

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Availability of data and materials

All data generated or analysed during this study are included in this published article and its Additional files 1, 2, 3, 4, and 5.
All contributing authors agreed consent for publication of the manuscript by Malaria journal.
The experiments were approved by the Experimental Animal Ethics Committee of Kyorin University School of Medicine, Tokyo.

Funding

This work was supported by a Grant-in-Aid for Scientific Research (C) from Japan Society for the Promotion of Science (JSPS) to MN (No. 15K08449). This work was also supported in part by a Grant-in-Aid for Scientific Research (A) from JSPS to KK (No. 26253025), JST/JICA Science and Technology Research Partnership for Sustainable Development (SATREPS) to KK (No. 10000284) and DKI (1400738), a Grant-in-Aid for Scientific Research (C) from the JSPS (No. 15K08451) to FK, Strategic International Research Cooperative Program (SICP) from the AMED to FK. We also acknowledge support from Program for Promotion of Basic and Applied Researches for Innovations in Bio-oriented Industry (BRAIN) and from Science and Technology Research Promotion Program for Agriculture, Forestry, Fisheries and Food Industry to KK and GSK Japan Research Grant 2016 for DKI.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated.
Anhänge

Additional files

Additional file 1. Fumarate cycle in Plasmodium falciparum. Fumarate, which is generated via purine biosynthesis, is converted into malate by fumarate hydratase (FH) [11]. Then, malate is converted to oxaloacetate (OOA) by malate:quinone oxidoreductase (MQO) [12]; the oxidation of malate to OOA generates ubiquinol (UQH2), which feeds the electron transport chain [11, 12]. Two of the eight mitochondrial TCA cycle enzymes, FH and MQO, may be essential for survival of asexual-blood-stage P. falciparum [9]. Note: MQO is conserved among all apicomplexan parasites, including all Cryptosporidium species [29].
Additional file 3. Generation of FH- and MQO-deficient Plasmodium berghei. SK-1 vector (A) and Sk-1-luc2 vector (B). Restriction sites of NheI and BglII restriction enzymes were shown. Schematic representation of gene-targeting vectors (A and B). Luciferase (luc2)-expressing cassette was introduced into target gene by double-crossover homologous recombination. Arrows (F1, F2, M1 and M2) denote primers specific for the 5′ and 3′ regions of the target gene (see Additional file 2). (A) Introduction of luc2-expressing cassette into the fh locus (PBANKA_082810) of P. berghei parasites. Proper integration was confirmed using primers specific for fh (WT, 3.0 kbp; Δfh, 6.8 kbp) for three cloned transfected parasites. (B) Introduction of luc2-expressing cassette into the mqo locus (PBANKA_111630) of P. berghei parasites. Proper integration was confirmed using primers specific for mqo (WT, 3.0 kbp; Δmqo, 7.0 kbp) for three cloned transfected parasites.
Additional file 4. Deficiency of FH and MQO has no effect on gametocyte production. Blood was obtained from infected mice showing 3% parasitaemia and cultured for 22 h under standardized in vitro culture conditions. Then, mature gametocytes and schizonts were collected by Nycodenz density-gradient centrifugation. (A and B) Expression of gametocyte-specific genes. mdv-1/peg3 [21] and g377 [22] were subjected to semi-quantitative RT–PCR using specific primers (see Additional files 2, 3). The hsp70 was used as a positive control. Samples treated with DNase-treated RNA template (hsp70 (-)) were used as a negative control that is the control of eventual DNA contamination of the RNA preparations. Experiments were performed in duplicate and representative data are shown. (C) Control and Δfh parasites-infected erythrocytes cultured for 22 h. (D) Control and Δmqo parasites-infected erythrocytes cultured for 22 h. White arrows indicate representative mature gametocytes. The scale bars indicate 20 μm. Note that sex-specific features such as nuclear enlargement, the distribution of pigment granules throughout the cytoplasm and enlargement of the cells are observed in both Δfh- and Δmqo-parasite cultures just the same as reported by Mons [23].
Additional file 5. Parasitaemia is suppressed by deficiency of parasite MQO but not FH. Bar graph of parasitaemia in Fig. 5A. Parasitaemia on days 3–7 post-infection were shown. Asterisks indicate a statistically significant difference (*, vs. control).
Literatur
2.
Zurück zum Zitat Voss TS, Bozdech Z, Bartfai R. Epigenetic memory takes center stage in the survival strategy of malaria parasites. Curr Opin Microbiol. 2014;20:88–95.CrossRefPubMed Voss TS, Bozdech Z, Bartfai R. Epigenetic memory takes center stage in the survival strategy of malaria parasites. Curr Opin Microbiol. 2014;20:88–95.CrossRefPubMed
3.
Zurück zum Zitat Lang-Unnasch N, Murphy AD. Metabolic changes of the malaria parasite during the transition from the human to the mosquito host. Annu Rev Microbiol. 1998;52:561–90.CrossRefPubMed Lang-Unnasch N, Murphy AD. Metabolic changes of the malaria parasite during the transition from the human to the mosquito host. Annu Rev Microbiol. 1998;52:561–90.CrossRefPubMed
4.
Zurück zum Zitat Gardner MJ, Hall N, Fung E, White O, Berriman M, Hyman RW, et al. Genome sequence of the human malaria parasite Plasmodium falciparum. Nature. 2002;419:498–511.CrossRefPubMed Gardner MJ, Hall N, Fung E, White O, Berriman M, Hyman RW, et al. Genome sequence of the human malaria parasite Plasmodium falciparum. Nature. 2002;419:498–511.CrossRefPubMed
5.
Zurück zum Zitat Bryant C, Voller A, Smith MJ. The incorporation of radioactivity from (14c)glucose into the soluble metabolic intermediates of malaria parasites. Am J Trop Med Hyg. 1964;13:515–9.CrossRefPubMed Bryant C, Voller A, Smith MJ. The incorporation of radioactivity from (14c)glucose into the soluble metabolic intermediates of malaria parasites. Am J Trop Med Hyg. 1964;13:515–9.CrossRefPubMed
6.
Zurück zum Zitat Scheibel LW, Pflaum WK. Cytochrome oxidase activity in platelet-free preparations of Plasmodium falciparum. J Parasitol. 1970;56:1054.CrossRefPubMed Scheibel LW, Pflaum WK. Cytochrome oxidase activity in platelet-free preparations of Plasmodium falciparum. J Parasitol. 1970;56:1054.CrossRefPubMed
7.
Zurück zum Zitat Hino A, Hirai M, Tanaka TQ, Watanabe Y, Matsuoka H, Kita K. Critical roles of the mitochondrial complex II in oocyst formation of rodent malaria parasite Plasmodium berghei. J Biochem. 2012;152:259–68.CrossRefPubMed Hino A, Hirai M, Tanaka TQ, Watanabe Y, Matsuoka H, Kita K. Critical roles of the mitochondrial complex II in oocyst formation of rodent malaria parasite Plasmodium berghei. J Biochem. 2012;152:259–68.CrossRefPubMed
8.
Zurück zum Zitat MacRae JI, Dixon MW, Dearnley MK, Chua HH, Chambers JM, Kenny S, et al. Mitochondrial metabolism of sexual and asexual blood stages of the malaria parasite Plasmodium falciparum. BMC Biol. 2013;11:67.CrossRefPubMedPubMedCentral MacRae JI, Dixon MW, Dearnley MK, Chua HH, Chambers JM, Kenny S, et al. Mitochondrial metabolism of sexual and asexual blood stages of the malaria parasite Plasmodium falciparum. BMC Biol. 2013;11:67.CrossRefPubMedPubMedCentral
9.
Zurück zum Zitat Ke H, Lewis IA, Morrisey JM, McLean KJ, Ganesan SM, Painter HJ, et al. Genetic investigation of tricarboxylic acid metabolism during the Plasmodium falciparum life cycle. Cell Rep. 2015;11:164–74.CrossRefPubMedPubMedCentral Ke H, Lewis IA, Morrisey JM, McLean KJ, Ganesan SM, Painter HJ, et al. Genetic investigation of tricarboxylic acid metabolism during the Plasmodium falciparum life cycle. Cell Rep. 2015;11:164–74.CrossRefPubMedPubMedCentral
10.
Zurück zum Zitat Raman J, Mehrotra S, Anand RP, Balaram H. Unique kinetic mechanism of Plasmodium falciparum adenylosuccinate synthetase. Mol Biochem Parasitol. 2004;138:1–8.CrossRefPubMed Raman J, Mehrotra S, Anand RP, Balaram H. Unique kinetic mechanism of Plasmodium falciparum adenylosuccinate synthetase. Mol Biochem Parasitol. 2004;138:1–8.CrossRefPubMed
11.
Zurück zum Zitat Bulusu V, Jayaraman V, Balaram H. Metabolic fate of fumarate, a side product of the purine salvage pathway in the intraerythrocytic stages of Plasmodium falciparum. J Biol Chem. 2011;286:9236–45.CrossRefPubMedPubMedCentral Bulusu V, Jayaraman V, Balaram H. Metabolic fate of fumarate, a side product of the purine salvage pathway in the intraerythrocytic stages of Plasmodium falciparum. J Biol Chem. 2011;286:9236–45.CrossRefPubMedPubMedCentral
12.
Zurück zum Zitat Uyemura SA, Luo S, Vieira M, Moreno SN, Docampo R. Oxidative phosphorylation and rotenone-insensitive malate- and NADH-quinone oxidoreductases in Plasmodium yoelii yoelii mitochondria in situ. J Biol Chem. 2004;279:385–93.CrossRefPubMed Uyemura SA, Luo S, Vieira M, Moreno SN, Docampo R. Oxidative phosphorylation and rotenone-insensitive malate- and NADH-quinone oxidoreductases in Plasmodium yoelii yoelii mitochondria in situ. J Biol Chem. 2004;279:385–93.CrossRefPubMed
13.
Zurück zum Zitat Storm J, Sethia S, Blackburn GJ, Chokkathukalam A, Watson DG, Breitling R, et al. Phosphoenolpyruvate carboxylase identified as a key enzyme in erythrocytic Plasmodium falciparum carbon metabolism. PLoS Pathog. 2014;10:e1003876.CrossRefPubMedPubMedCentral Storm J, Sethia S, Blackburn GJ, Chokkathukalam A, Watson DG, Breitling R, et al. Phosphoenolpyruvate carboxylase identified as a key enzyme in erythrocytic Plasmodium falciparum carbon metabolism. PLoS Pathog. 2014;10:e1003876.CrossRefPubMedPubMedCentral
14.
Zurück zum Zitat Srivastava A, Creek DJ, Evans KJ, De Souza D, Schofield L, Muller S, et al. Host reticulocytes provide metabolic reservoirs that can be exploited by malaria parasites. PLoS Pathog. 2015;11:e1004882.CrossRefPubMedPubMedCentral Srivastava A, Creek DJ, Evans KJ, De Souza D, Schofield L, Muller S, et al. Host reticulocytes provide metabolic reservoirs that can be exploited by malaria parasites. PLoS Pathog. 2015;11:e1004882.CrossRefPubMedPubMedCentral
15.
Zurück zum Zitat Niikura M, Inoue S, Mineo S, Yamada Y, Kaneko I, Iwanaga S, et al. Experimental cerebral malaria is suppressed by disruption of nucleoside transporter 1 but not purine nucleoside phosphorylase. Biochem Biophys Res Commun. 2013;432:504–8.CrossRefPubMed Niikura M, Inoue S, Mineo S, Yamada Y, Kaneko I, Iwanaga S, et al. Experimental cerebral malaria is suppressed by disruption of nucleoside transporter 1 but not purine nucleoside phosphorylase. Biochem Biophys Res Commun. 2013;432:504–8.CrossRefPubMed
16.
Zurück zum Zitat Ecker A, Moon R, Sinden RE, Billker O. Generation of gene targeting constructs for Plasmodium berghei by a PCR-based method amenable to high throughput applications. Mol Biochem Parasitol. 2006;145:265–8.CrossRefPubMed Ecker A, Moon R, Sinden RE, Billker O. Generation of gene targeting constructs for Plasmodium berghei by a PCR-based method amenable to high throughput applications. Mol Biochem Parasitol. 2006;145:265–8.CrossRefPubMed
17.
Zurück zum Zitat Janse CJ, Franke-Fayard B, Mair GR, Ramesar J, Thiel C, Engelmann S, et al. High efficiency transfection of Plasmodium berghei facilitates novel selection procedures. Mol Biochem Parasitol. 2006;145:60–70.CrossRefPubMed Janse CJ, Franke-Fayard B, Mair GR, Ramesar J, Thiel C, Engelmann S, et al. High efficiency transfection of Plasmodium berghei facilitates novel selection procedures. Mol Biochem Parasitol. 2006;145:60–70.CrossRefPubMed
18.
Zurück zum Zitat Janse CJ, Ramesar J, Waters AP. High-efficiency transfection and drug selection of genetically transformed blood stages of the rodent malaria parasite Plasmodium berghei. Nat Protoc. 2006;1:346–56.CrossRefPubMed Janse CJ, Ramesar J, Waters AP. High-efficiency transfection and drug selection of genetically transformed blood stages of the rodent malaria parasite Plasmodium berghei. Nat Protoc. 2006;1:346–56.CrossRefPubMed
19.
Zurück zum Zitat Kawahara K, Mogi T, Tanaka TQ, Hata M, Miyoshi H, Kita K. Mitochondrial dehydrogenases in the aerobic respiratory chain of the rodent malaria parasite Plasmodium yoelii yoelii. J Biochem. 2009;145:229–37.CrossRefPubMed Kawahara K, Mogi T, Tanaka TQ, Hata M, Miyoshi H, Kita K. Mitochondrial dehydrogenases in the aerobic respiratory chain of the rodent malaria parasite Plasmodium yoelii yoelii. J Biochem. 2009;145:229–37.CrossRefPubMed
20.
Zurück zum Zitat Thumwood CM, Hunt NH, Clark IA, Cowden WB. Breakdown of the blood–brain barrier in murine cerebral malaria. Parasitology. 1988;96(Pt 3):579–89.CrossRefPubMed Thumwood CM, Hunt NH, Clark IA, Cowden WB. Breakdown of the blood–brain barrier in murine cerebral malaria. Parasitology. 1988;96(Pt 3):579–89.CrossRefPubMed
21.
Zurück zum Zitat Ponzi M, Siden-Kiamos I, Bertuccini L, Curra C, Kroeze H, Camarda G, et al. Egress of Plasmodium berghei gametes from their host erythrocyte is mediated by the MDV-1/PEG3 protein. Cell Microbiol. 2009;11:1272–88.CrossRefPubMed Ponzi M, Siden-Kiamos I, Bertuccini L, Curra C, Kroeze H, Camarda G, et al. Egress of Plasmodium berghei gametes from their host erythrocyte is mediated by the MDV-1/PEG3 protein. Cell Microbiol. 2009;11:1272–88.CrossRefPubMed
22.
Zurück zum Zitat Janse CJ, Haghparast A, Speranca MA, Ramesar J, Kroeze H, del Portillo HA, et al. Malaria parasites lacking eef1a have a normal S/M phase yet grow more slowly due to a longer G1 phase. Mol Microbiol. 2003;50:1539–51.CrossRefPubMed Janse CJ, Haghparast A, Speranca MA, Ramesar J, Kroeze H, del Portillo HA, et al. Malaria parasites lacking eef1a have a normal S/M phase yet grow more slowly due to a longer G1 phase. Mol Microbiol. 2003;50:1539–51.CrossRefPubMed
23.
Zurück zum Zitat Mons B. Intra erythrocytic differentiation of Plasmodium berghei. Acta Leiden. 1986;54:1–124.PubMed Mons B. Intra erythrocytic differentiation of Plasmodium berghei. Acta Leiden. 1986;54:1–124.PubMed
24.
Zurück zum Zitat Lelievre J, Almela MJ, Lozano S, Miguel C, Franco V, Leroy D, Herreros E. Activity of clinically relevant antimalarial drugs on Plasmodium falciparum mature gametocytes in an ATP bioluminescence “transmission blocking” assay. PLoS ONE. 2012;7:e35019.CrossRefPubMedPubMedCentral Lelievre J, Almela MJ, Lozano S, Miguel C, Franco V, Leroy D, Herreros E. Activity of clinically relevant antimalarial drugs on Plasmodium falciparum mature gametocytes in an ATP bioluminescence “transmission blocking” assay. PLoS ONE. 2012;7:e35019.CrossRefPubMedPubMedCentral
25.
Zurück zum Zitat Nozawa A, Fujimoto R, Matsuoka H, Tsuboi T, Tozawa Y. Cell-free synthesis, reconstitution, and characterization of a mitochondrial dicarboxylate–tricarboxylate carrier of Plasmodium falciparum. Biochem Biophys Res Commun. 2011;414:612–7.CrossRefPubMed Nozawa A, Fujimoto R, Matsuoka H, Tsuboi T, Tozawa Y. Cell-free synthesis, reconstitution, and characterization of a mitochondrial dicarboxylate–tricarboxylate carrier of Plasmodium falciparum. Biochem Biophys Res Commun. 2011;414:612–7.CrossRefPubMed
26.
Zurück zum Zitat Sturm A, Mollard V, Cozijnsen A, Goodman CD, McFadden GI. Mitochondrial ATP synthase is dispensable in blood-stage Plasmodium berghei rodent malaria but essential in the mosquito phase. Proc Natl Acad Sci USA. 2015;112:10216–23.CrossRefPubMedPubMedCentral Sturm A, Mollard V, Cozijnsen A, Goodman CD, McFadden GI. Mitochondrial ATP synthase is dispensable in blood-stage Plasmodium berghei rodent malaria but essential in the mosquito phase. Proc Natl Acad Sci USA. 2015;112:10216–23.CrossRefPubMedPubMedCentral
27.
Zurück zum Zitat Akiba T, Hiraga K, Tuboi S. Intracellular distribution of fumarase in various animals. J Biochem. 1984;96:189–95.CrossRefPubMed Akiba T, Hiraga K, Tuboi S. Intracellular distribution of fumarase in various animals. J Biochem. 1984;96:189–95.CrossRefPubMed
28.
Zurück zum Zitat Tanaka KR, Valentine WN. Fumarase activity of human leukocytes and erythrocytes. Blood. 1961;17:328–33.PubMed Tanaka KR, Valentine WN. Fumarase activity of human leukocytes and erythrocytes. Blood. 1961;17:328–33.PubMed
29.
Zurück zum Zitat Mogi T, Kita K. Diversity in mitochondrial metabolic pathways in parasitic protists Plasmodium and Cryptosporidium. Parasitol Int. 2010;59:305–12.CrossRefPubMed Mogi T, Kita K. Diversity in mitochondrial metabolic pathways in parasitic protists Plasmodium and Cryptosporidium. Parasitol Int. 2010;59:305–12.CrossRefPubMed
30.
Zurück zum Zitat Gottlieb E, Armour SM, Harris MH, Thompson CB. Mitochondrial membrane potential regulates matrix configuration and cytochrome c release during apoptosis. Cell Death Differ. 2003;10:709–17.CrossRefPubMed Gottlieb E, Armour SM, Harris MH, Thompson CB. Mitochondrial membrane potential regulates matrix configuration and cytochrome c release during apoptosis. Cell Death Differ. 2003;10:709–17.CrossRefPubMed
31.
Zurück zum Zitat Nixon GL, Moss DM, Shone AE, Lalloo DG, Fisher N, O’Neill PM, et al. Antimalarial pharmacology and therapeutics of atovaquone. J Antimicrob Chemother. 2013;68:977–85.CrossRefPubMedPubMedCentral Nixon GL, Moss DM, Shone AE, Lalloo DG, Fisher N, O’Neill PM, et al. Antimalarial pharmacology and therapeutics of atovaquone. J Antimicrob Chemother. 2013;68:977–85.CrossRefPubMedPubMedCentral
32.
Zurück zum Zitat Goodman CD, Siregar JE, Mollard V, Vega-Rodriguez J, Syafruddin D, Matsuoka H, et al. Parasites resistant to the antimalarial atovaquone fail to transmit by mosquitoes. Science. 2016;352:349–53.CrossRefPubMedPubMedCentral Goodman CD, Siregar JE, Mollard V, Vega-Rodriguez J, Syafruddin D, Matsuoka H, et al. Parasites resistant to the antimalarial atovaquone fail to transmit by mosquitoes. Science. 2016;352:349–53.CrossRefPubMedPubMedCentral
Metadaten
Titel
Suppression of experimental cerebral malaria by disruption of malate:quinone oxidoreductase
verfasst von
Mamoru Niikura
Keisuke Komatsuya
Shin-Ichi Inoue
Risa Matsuda
Hiroko Asahi
Daniel Ken Inaoka
Kiyoshi Kita
Fumie Kobayashi
Publikationsdatum
01.12.2017
Verlag
BioMed Central
Erschienen in
Malaria Journal / Ausgabe 1/2017
Elektronische ISSN: 1475-2875
DOI
https://doi.org/10.1186/s12936-017-1898-5

Weitere Artikel der Ausgabe 1/2017

Malaria Journal 1/2017 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Notfall-TEP der Hüfte ist auch bei 90-Jährigen machbar

26.04.2024 Hüft-TEP Nachrichten

Ob bei einer Notfalloperation nach Schenkelhalsfraktur eine Hemiarthroplastik oder eine totale Endoprothese (TEP) eingebaut wird, sollte nicht allein vom Alter der Patientinnen und Patienten abhängen. Auch über 90-Jährige können von der TEP profitieren.

Niedriger diastolischer Blutdruck erhöht Risiko für schwere kardiovaskuläre Komplikationen

25.04.2024 Hypotonie Nachrichten

Wenn unter einer medikamentösen Hochdrucktherapie der diastolische Blutdruck in den Keller geht, steigt das Risiko für schwere kardiovaskuläre Ereignisse: Darauf deutet eine Sekundäranalyse der SPRINT-Studie hin.

Bei schweren Reaktionen auf Insektenstiche empfiehlt sich eine spezifische Immuntherapie

Insektenstiche sind bei Erwachsenen die häufigsten Auslöser einer Anaphylaxie. Einen wirksamen Schutz vor schweren anaphylaktischen Reaktionen bietet die allergenspezifische Immuntherapie. Jedoch kommt sie noch viel zu selten zum Einsatz.

Therapiestart mit Blutdrucksenkern erhöht Frakturrisiko

25.04.2024 Hypertonie Nachrichten

Beginnen ältere Männer im Pflegeheim eine Antihypertensiva-Therapie, dann ist die Frakturrate in den folgenden 30 Tagen mehr als verdoppelt. Besonders häufig stürzen Demenzkranke und Männer, die erstmals Blutdrucksenker nehmen. Dafür spricht eine Analyse unter US-Veteranen.

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.