Abstract
Pediatric heart failure remains poorly understood, distinct in many aspects from adult heart failure. Limited data point to roles of altered mitochondrial functioning and, in particular, changes in mitochondrial lipids, especially cardiolipin. Barth syndrome is a mitochondrial disorder caused by tafazzin mutations that lead to abnormal cardiolipin profiles. Patients are afflicted by cardiomyopathy, skeletal myopathy, neutropenia, and growth delay. A mouse model of Barth syndrome was developed a decade ago, which relies on a doxycycline-inducible short hairpin RNA to knock down expression of tafazzin mRNA (TAZKD). Our objective was to review published data from the TAZKD mouse to determine its contributions to our pathogenetic understanding of, and potential treatment strategies for, Barth syndrome. In regard to the clinical syndrome, the reported physiological, biochemical, and ultrastructural abnormalities of the mouse model mirror those in Barth patients. Using this model, the peroxisome proliferator-activated receptor pan-agonist bezafibrate has been suggested as potential therapy because it ameliorated the cardiomyopathy in TAZKD mice, while increasing mitochondrial biogenesis. A clinical trial is now underway to test bezafibrate in Barth syndrome patients. Thus the TAZKD mouse model of Barth syndrome has led to important insights into disease pathogenesis and therapeutic targets, which can potentially translate to pediatric heart failure.
Keywords: Barth syndrome, mouse model, pediatric heart failure, tafazzin
INTRODUCTION
Pediatric heart failure is a poorly understood condition with significant morbidity and mortality. Pathophysiological etiologies of heart failure have been difficult to define, but there is significant overlap between pediatric heart failure, congenital heart disease, cardiomyopathy, rhythm abnormalities, and acquired conditions, such as myocarditis (19, 60, 67, 68, 91). Etiology and symptoms differ by age group, and cellular mechanisms of the young heart are unique from adult hearts, thus requiring a better understanding of the pathophysiology of heart failure in children (60, 100). Therapeutic options for pediatric heart failure remain poorly developed due to the challenges of extrapolating adult data to children (19, 66, 100).
Impaired mitochondrial function is an important contributor to the impaired cardiac function in heart failure (28, 55, 57). Alterations in mitochondrial membrane lipid composition may contribute to abnormal mitochondrial function (25). Cardiolipin (CL) is a major cardiac phospholipid found in the inner mitochondrial membrane with essential roles in respiratory chain, metabolite carriers, and mitochondrial metabolism (22). Loss of CL leads to the generation of excessive reactive oxygen species (ROS) as a result of inefficient electron transport chain complexes (20). CL content has been shown to be significantly altered in pediatric heart failure, indicating a role of CL biosynthesis and regulation in the pathogenesis of heart failure (14, 80). Most famously, alterations in CL biochemistry are implicated in the pathogenesis of Barth syndrome (BTHS) and Barth cardiomyopathy (25, 46). Barth syndrome (MIM 302060) is a rare X-linked mitochondrial disorder in which patients are afflicted by cardiomyopathy, skeletal myopathy, neutropenia, and growth delay. Mutations in a nuclear-encoded mitochondrial phospholipid transacylase (tafazzin) lead to CL deficiency and mitochondrial dysfunction. Tafazzin catalyzes CL remodeling reactions at the final stage of CL biosynthesis, producing “mature” CL, which is critical to the structural integrity and diverse functions of mitochondria (27, 34, 61, 72, 77). A mouse model of Barth syndrome was developed almost 10 yr ago. This model relies on a doxycycline-inducible short hairpin RNA (shRNA) to knock down expression of tafazzin mRNA (TAZKD).
The goals of any model include a deeper understanding of disease pathogenesis and translational insights into potential therapies. The purpose of this review is to summarize the published data from the TAZKD mouse to determine its contributions to our pathogenetic understanding of, and potential treatment strategies toward, Barth syndrome. We focused on the following questions: 1) Does the mouse mimic the clinical syndrome? 2) Has this model led to a deeper understanding of disease pathogenesis? 3) Have potential therapies been suggested by this model?
METHODS
We searched PubMed and Medline using the terms “mouse,” “tafazzin,” and “knockdown,” as well as “TAZKD.” We selected all published articles that used the TAZKD mouse model to generate data. Unpublished data, data from abstracts only, or data presented at conferences but not published were not included in this analysis. We also reviewed selected literature of clinical data on Barth syndrome patients.
RESULTS
TAZKD Model: Genetic Engineering, Knockdown Induction, and Differences in TAZKD Induction
Transgenic mice were originally generated at TaconicArtemis (Köln, Germany) under contract from the Barth Syndrome Foundation; these mice are now available through The Jackson Laboratory (stock no. 014648; Bar Harbor, ME). These TAZKD mice have been previously described in detail (1, 79) on the basis of a tight inducible shRNA expression system (74) (Fig. 1). TAZKD transgenic mice used in studies are the heterozygous offspring of heterozygous (male)-wild-type (female) C57Bl/6 crosses, as per TaconicArtemis’ protocol [Phoon et al. (58)]. Doxycycline was administered in drinking water or in chow to induce knockdown of tafazzin. An advantage of this system is that tafazzin knockdown can be “turned on” as desired.
Fig. 1.
A: inserted by recombinase-mediated cassette exchange, the F3-FRT cassette contains the tafazzin (taz)-specific short hairpin RNA (shRNA) H1-tet-On promoter, a tet repressor (tetR), and a neomycin selection gene. B: the tetO-shRNA promoter is inactivated by tetR until doxycycline (DOXY) is added to the system, inactivating tetR, and allowing transcription of taz-specific shRNA. The system is reversible. Different DOXY tafazzin knockdown (TAZKD) induction protocols are listed. Knockdown of tafazzin will lead to rapid decreases in taz mRNA, but the time course of decreases in tafazzin protein, changes in the cardiolipin (CL) profiles, changes in mitochondrial ultrastructure and function, and changes at the cellular, tissue, and organ levels have not been precisely determined. L4CL, tetralinoleoyl cardiolipin; MLCL, monolyocardiolipin; SA, Splicing acceptor signal. [Adapted from Phoon et al. (58). Used with permission.]
The initial characterization studies notably used three different doxycycline induction, taz-knockdown protocols (1, 58, 79) (Fig. 1). In the Phoon et al. (58) study, doxycycline was administered in the drinking water [2 mg/mL and sucrose 10% (for palatability)] to wild-type female mice on a continuous basis, to make sure doxycycline remained in the maternal system and the embryonic transgene would, therefore, be induced throughout gestation. This was the same induction protocol as in the original TaconicArtemis protocol (74). However, because the males carry the transgene as heterozygotes, mating pairs were fed regular water without doxycycline to reduce the possibility of male sterility, which is a known adverse effect in a Drosophila model of tafazzin deficiency (50) and a more recently developed tafazzin knockout (TAZKO) mouse model (62). Doxycycline was reintroduced into the drinking water (2 mg/mL and sucrose 10%) of the females on the day following overnight mating (day of plug). Of note, this doxycycline dosing differed from the first two published investigations on the TAZKD mouse model [625 mg/kg chow (1); 200 mg/kg chow (79)]. It was estimated in the Phoon study (58) that TAZKD mice ingested 3–10 times the doxycycline used in studies by Acehan et al. (1) and Soustek et al. (79), because mice ingest, on average, 15 g/100 g body wt in food and 15 mL/100 g body wt in water daily (https://www.lvma.org/mouse.html).
Administration of doxycycline induces the shRNA expression system of the TAZKD model to decrease tafazzin mRNA and protein levels (Fig. 1). Doxycycline dosage has been shown to influence the speed at which the knockdown achieves steady-state levels (74). Various studies have differed in the timing of doxycycline administration as opposed to starting administration at the start of gestation [in drinking water from 3 mo of age (44); 625 mg/kg in chow from 2 mo of age (90)]. Still other studies combined induction methods, starting with 625 mg/kg chow from the start of gestation and switching to drinking water [1 mg/mL (32); 0.5 mg/mL (71)] at 3 mo of age. Data show that taz knockdown is highly efficient, with >90–95% knockdown of mRNA and protein at 2 mo of age when induced prenatally (1, 79) in heart and to somewhat varying degrees in other organs, such as liver, skeletal muscle, and brain. In embryos, 80% knockdown of tafazzin mRNA could be achieved within 3 days (58).
As will be seen, the different doxycycline dosing regimens and taz-knockdown induction protocols are important in the interpretation of published data in this mouse model. We note that, while mRNA may be knocked down rapidly, tafazzin levels and/or CL may take additional time to decrease, particularly since the half-life of CL is quite long (48, 98, 102); however, the timeline required to achieve tafazzin protein deficiency (to very low, steady-state levels) has not been carefully determined due to the lack of a good antibody. Once CL profiles have been altered, it may take time for a cellular and tissue/organ phenotype to develop (Fig. 1). The kinetics of sufficient protein knockdown indicate that the results of the model are based on the minimum time that knockdown is induced, which limits conclusions from some studies that did not induce knockdown for at least 3 mo. This may partially explain the difference in results across investigations based on the dosage and timing of induction. Some studies may not have allowed enough time for tafazzin protein depletion or for significant CL abnormalities to occur and, therefore, did not observe the same abnormalities that have been described in other investigations of this model (43, 75, 90).
A recent study by Moullan et al. (54) issued a caution on the potential confounding effects of doxycycline on experimental results in biomedical research. In addition, the high concentration of sucrose in the drinking water for some TAZKD protocols may further confound the interpretation of the results.
Phenotype
Cardiac abnormalities.
In the TAZKD mouse, two distinct cardiac phenotypes created by two different doxycycline induction protocols have been reported: 1) perinatal lethality, due to severe diastolic dysfunction associated with left ventricular (LV) noncompaction (LVNC) (58), and 2) a mild, adult-onset dilated cardiomyopathy (DCM) (1, 79). With tafazzin knockdown induced by the higher dose doxycycline-in-drinking-water protocol, the prenatal TAZKD model demonstrated complete lethality by the neonatal period and significant loss of TAZKD embryos at embryonic days 12.5–14.5 (58). Echocardiography revealed diastolic cardiac dysfunction, but with preserved systolic function, in embryonic hearts. Abnormalities in mitochondrial ultrastructure and sarcomeric organization were seen, including vacuolated cristae, a reduction in total mitochondrial area density, a reduction in mitochondrial size, and lower cristae density. Also, newborn TAZKD mice demonstrated a disruption of the normal parallel orientation between mitochondria and sarcomeres. Overall, these abnormalities indicate abnormal mitochondria and are consistent with delayed cardiomyocyte differentiation. Histology of the embryonic and newborn TAZKD hearts demonstrate altered cellular proliferation in association with the myocardial hypertrabeculation noncompaction, suggesting abnormalities in myocardial formation and patterning (58).
In contrast, in TAZKD mice induced with the lower doses of doxycycline in chow, there has not been evidence of significant defects in cardiac function in 2-mo-old TAZ knockdown mice under basal conditions. Cardiac dysfunction was not reported to develop until 8 mo of age (1, 21), although a more recent report suggested cardiac dysfunction by 5 mo of age (88). The dramatic differences in the survival and cardiac phenotype have been attributed to the differences in doxycycline dosing (74). In an adrenergic stress model of the TAZKD mouse, induced by chronic isoproterenol infusion, the cardiac phenotype was rapidly exacerbated with DCM and LV systolic dysfunction observed by 4.5 mo of age, compared with 7 mo of age without β-adrenergic stress (32). Thus the cardiomyopathy induced by this doxycycline protocol becomes apparent by echocardiography between 5 and 8 mo of age.
The cardiac dysfunction observed in the TAZKD mouse includes LV dilation, loss of LV muscle mass, and decrease in LV ejection force (1, 79). In yet another TAZKD mouse study, hypertrophic cardiomyopathy (HCM) instead of DCM was seen (37). Cardiac muscle of the TAZKD model demonstrates disturbed sarcomere organization, mitochondrial proliferation, myofibrillar disarray, mitophagy, and mitochondria-associated membrane abnormalities (1, 79).
“Acquired” tafazzin deficiency can be studied as well, to circumvent any developmental abnormalities that might arise and to mimic adult-onset cardiac diseases. Kim and coworkers (43) induced tafazzin knockdown at weaning and studied 8- to 12-wk-old TAZKD mice; therefore, the period of doxycycline induction spanned 4–8 wk only. Total CL was reduced by 50%, with a dramatic, 25-fold increase in monolysocardiolipin (MLCL). The investigators isolated subsarcolemmal and interfibrillar cardiac mitochondria and identified significant mitochondrial abnormalities, including changes in mitochondrial ultrastructure and decomposition of respiratory supercomplexes (with a concomitant increase in supercomplex I), but oxidative phosphorylation (OXPHOS) was unchanged in TAZKD mice. The investigators speculated that OXPHOS was compensated by an increase in supercomplex I (43). Because altered CL profiles have been implicated in ischemia-reperfusion injury, Szczepanek and coworkers (90) investigated how acquired tafazzin deficiency would impact mitochondrial function and response to cardiac injury (Langendorff model: 25 min ischemia + 60 min reperfusion). The TAZKD mouse was induced postnatally, from 2 to 4 mo of age, such that tafazzin deficiency would be “acquired” in young adult mice. Although TAZKD hearts showed increased ROS and increased susceptibility to mitochondrial permeability transition pore opening at baseline, the rate of oxygen consumption was unchanged, as was the response (LV infarct size) to ischemia-reperfusion injury. It should be noted that TAZKD mice were induced with doxycycline for only 2 mo, as reflected by the only 40% decrease in CL.
Clearly, the TAZKD mouse exhibits a cardiomyopathy, as does the human Barth syndrome. It is not clear at this time why some investigators have reported a DCM while others have found an HCM in the TAZKD mouse. The prenatal lethality of the TAZKD mouse model (58) also mirrors the male fetal loss and stillbirth that results from the fetal cardiomyopathy associated with Barth syndrome (86). However, because cardiomyopathy typically presents very early in Barth patients (most often within the first year), the delayed presentation of cardiac defects in the TAZKD mouse model has led some to criticize the model’s representation of the human condition (43, 90). Since human Barth patients will exhibit stabilization of cardiac function at older ages, an alternative explanation is that the TAZKD mouse does not exhibit an infantile cardiomyopathy/heart failure phenotype. Endocardial fibroelastosis has not been described in the TAZKD mouse. Moreover, TAZKD mice have not demonstrated any significant conduction abnormalities or ventricular arrhythmias, in contrast to findings in human Barth patients (83) (summarized in Tables 1 and 2).
Table 1.
Cardiac abnormalities
Human Barth Patients | TAZKD Mouse Model | |
---|---|---|
Diagnosed in young patients (<5 yr of age and infants): Fetal cardiomyopathy and heart failure Dilated cardiomyopathy (DCM) Left ventricular noncompaction cardiomyopathy (LVNC) Hypertrophic cardiomyopathy (HCM) |
Diagnosed in young mice (prenatal or <2 mo of age): Pre-/perinatal LVNC cardiomyopathy |
|
Variable course and severity | Ventricular arrhythmia Endocardial fibroelastosis Sudden cardiac death |
4 mo of age No change in heart contractile recovery after ischemia-reperfusion Chronic isoproterenol treatment (age 2.5–3 mo) exacerbated the cardiac dysfunction at age 4.5 mo LVFS↓ LVEF↓ IVSd↓ 6–7 mo of age Hypertrophic cardiomyopathy 7–8 mo of age Dilated cardiomyopathy LVFS↓ LVEF↓ LV dilation |
Cardiac response to exercise | Blunted contractile function Impaired cardiac reserve ↓Peak O2 consumption and work rate ↓PCr recovery Response to endurance training: ↑Peak O2 consumption (modest) No changes in mean heart rate, cardiac function Improved quality of life scores, symptoms |
↓Peak O2 consumption Response to endurance training: Cardiac hypertrophy ↓Complex I, III, and IV enzymatic activities at baseline, with ↑complex III activity with endurance training |
Cardiomyocytes | Irregular sarcomere organization Abnormal mitochondrial structure |
Abnormal sarcomere organization Abnormal mitochondrial ultrastructure |
References | 5, 8, 13, 16, 21, 26, 38, 56, 64, 81–83, 86, 99 | 1, 21, 32, 37, 58, 59, 71, 78, 79, 90 |
↑, Increase; ↓, decrease; IVSd, interventricular septum thickness at end-diastole; LVEF, left ventricular ejection fraction; LVFS, left ventricular fractional shortening; TAZKD, knock down expression of tafazzin mRNA.
Table 2.
Phenotype and pathophysiology
Human Barth Patients | TAZKD Mouse Model | |
---|---|---|
Fetal loss | 100% pre-/perinatal lethality | |
Skeletal | Delayed motor milestones Lipid storage myopathy Proximal myopathy Abnormal fatigability Exercise intolerance Impaired balance and motion reaction time ↓O2 utilization and extraction ↓Phosphocreatine/ATP ratio ↓PCr recovery rate Blunted fat oxidation |
Soleus muscle weakness Decreased exercise capacity Impaired O2 utilization |
Growth | Abnormal growth trajectories Constitutional growth delay with delayed bone age Failure to thrive Need for assistive feeding or nutritional support |
Lagging growth Decreased body weight Lower fat and lean mass |
Hematological | Variable neutropenia | Low neutrophil count |
Neurological | Cognitive difficulties Psychological and sensory abnormalities |
Significant memory deficiency |
References | 5, 8, 12, 16, 29, 36, 52, 63–65, 81, 82, 85–87 | 1, 17, 18, 33, 37, 58, 59, 79, 82 |
↓, Decrease; TAZKD, knock down expression of tafazzin mRNA.
Skeletal myopathy.
The TAZKD mouse also exhibits significant reduction in skeletal muscle function (79). There is an observed reduction in muscle strength, with significant decreased soleus muscle contractility and diminished exercise capacity in the mouse model (59, 79). The TAZKD mouse also exhibits impaired oxygen utilization at high workload, consistent with findings in human Barth patients (59, 82). These functional deficiencies are accompanied by ultrastructural abnormalities, including aggregation, disrupted cristae, size, and shape variations, in mitochondrial and muscle fiber (1).
The skeletal and mitochondrial myopathy associated with exercise intolerance present in the TAZKD mouse is representative of the human condition of Barth syndrome (29, 82) (Table 2).
Growth delay.
The initial characterization of the TAZKD mouse showed normal growth through the first 2 mo of age, but lagging growth by 8 mo (1). TAZKD mice typically weigh less than their wild-type littermates due to hypermetabolism (18). On average, TAZKD mice weigh 16% less than their wild-type littermates (33). This weight difference has been attributed to lower fat and lean mass in TAZKD mice due to elevated energy expenditure and activity levels (37). Growth delay appears similar to human BTHS (Table 2).
Additional characteristics.
In a single study, low neutrophil count was observed in 50% of TAZKD mice, although the data were not explicitly reported (79). However, the investigators noted that this result is complicated by the use of doxycycline, which exerts reductive effects on neutrophil counts. Significant memory deficiencies have also recently been identified in the TAZKD mouse (17). However, the TAZKD data are scant for a thorough comparison with human BTHS (Table 2).
Disease Pathogenesis
See Table 3.
Table 3.
Biochemistry
Human Barth Patients | TAZKD Mouse Model |
---|---|
Abnormal molecular composition of cardiolipin, phosphatidylcholine, and phosphatidylethanolamine | Taz mRNA↓ Taz protein↓ Differential expression of TMEM65 |
TLCL↓ | TLCL↓ |
↑MLCL:CL ratio >0.3 is threshold for diagnosis |
MLCL/CL↑↑ Plasmalogen↓ OCR↓ Reduced components of mitochondrial respiratory chain (especially I, III, IV) |
↓Cytochromes | |
Decreased mitochondrial respiratory chain complex activity in skeletal muscle, cultured fibroblasts, and lymphoblasts | |
Decrease in supercomplexes | Decrease in supercomplexes |
Cardiac-specific loss of succinate dehydrogenase | Cardiac-specific loss of succinate dehydrogenase |
3-Methylglutaconic aciduria ↓Arginine level |
↓Mitochondria bound myoglobin |
↓Prealbumin level | |
↓Serum cholesterol | |
ROS↑ | ROS↑ |
Lactic acidemia | ↓Fatty acid oxidation enzymes |
Blunted fat oxidation and elevated glucose metabolism upon exercise | ↑Cytosolic glycolysis Abnormal mitochondria Mitophagy Paracrystalline structures Vacuolated cristae |
Structural and functional abnormalities of mitochondria Inner membrane adhesions Disrupted cristae | |
Refs. 2, 3, 5–7, 9, 12, 15, 21, 24, 30, 39, 40, 47, 53, 64, 70, 73, 81, 93, 95, 97, 99, 101, 103 | Refs. 1, 21, 31, 37, 42, 44, 58, 59, 71, 78, 79, 89, 90, 101 |
↑, Increase; ↓, decrease; CL, cardiolipin; MLCL; monolysocardiolipin; TAZKD, knock down expression of tafazzin mRNA; TLCL, tetralinoleoyl cardiolipin; TMEM65, transmembrane protein 65.
Phospholipid composition, profiles, and alterations.
On induction, the knock down of tafazzin has significant effects on the overall lipid composition of the mitochondrial membrane. The primary finding is that tafazzin knockdown causes tetralinoleoyl CL deficiency with MLCL accumulation, specifically in cardiac and skeletal muscle (1, 21, 90). Notably, different tissues respond differently to tafazzin deficiency, because they exhibit different CL profiles, with less tetralinoleoyl CL in organs such as kidney, liver, and brain. Thus levels of tetralinoleoyl CL in the liver, brain, and white (but not brown) adipose tissue of TAZKD mice are similar to wild-type control levels (17, 18, 21). However, MLCL increases in all studied tissues in the TAZKD mouse, thereby increasing the MLCL-to-CL ratio to varying degrees. The increased MLCL-to-CL ratio has been observed in both adult and embryonic mouse hearts (1, 58, 79). The mouse demonstrates accumulation of MLCL and dilysocardiolipin, altered distribution of acyl chains in choline and ethanolamine glycerophospholipids, and dysregulated generation of potent oxidized lipid metabolites critical for hemodynamic function (42). It has also been found that the TAZKD mouse exhibits decreased plasmalogen in the heart (44). Xu and coworkers (101) found that CL is tightly bound to proteins, while MLCL is not, in the TAZKD mouse tissues as well as human lymphoblasts.
Respiratory chain complexes.
The disruption of the composition of the inner mitochondrial membrane as a result of CL abnormalities destabilizes respiratory chain complexes by disrupting supercomplex assembly (21). Organization of both respiratory chain complexes and supercomplexes is disrupted in TAZKD mitochondria, affecting electron flux, proton gradient, and ATP synthesis (31). The disruption of supercomplex formation is also tissue specific and mirrors the abnormalities in CL profiles; therefore, while heart, skeletal muscle, and brown adipose tissue exhibit destabilization of supercomplexes (17, 18, 21, 31, 101), liver, brain, and white adipose tissue are not substantially affected (17, 18). Once again, it is apparent that heart (and skeletal) muscle is most severely affected. Seahorse extracellular flux analysis shows that oxygen consumption rate in TAZKD mitochondria is significantly reduced compared with controls (21). There is a decrease in the abundance of large protein assemblies in the TAZKD mouse, including respiratory supercomplexes, ATP synthase oligomers, and complexes containing the ADP-ATP carrier (101). TAZKD mouse hearts have reduced components of the mitochondrial respiratory chain (31). While some results have shown a reduction in complex III activity with complex IV activity not affected in cardiac mitochondria (21, 42, 59), another study found that there is a reduction in complex I, III, and IV activity in cardiac tissue, and that exercise training improves complex III activity, suggesting different metabolic abnormalities under different conditions of exercise (78). Interestingly, while respiratory chain complexes are significantly affected in TAZKD mouse mitochondria, the ATPase complex is not (21).
Metabolism.
Proteomic analysis of TAZKD mouse hearts demonstrated marked metabolic remodeling, including disruptions in fatty acid oxidation enzymes and reduced mitochondria-bound myoglobin, potentially disrupting intracellular oxygen delivery to the OXPHOS system (31). Maximal capacity of OXPHOS is lowered in cardiac mitochondria from the TAZKD mouse (37); consequently, CL-deficient cardiomyocytes have an increased reliance on cytosolic glycolysis (59). The mouse demonstrates a downregulation of genes related to O2/CO2 exchange and aerobic metabolism (71), and a shift in preference for glutamate (amino acid) has been found to stimulate oxidation over fatty acid metabolism (42, 71). Additionally, the formation of ROS, specifically hydrogen peroxide (H2O2), is significantly increased in the TAZKD mouse (21, 59, 90).
Mitochondrial ultrastructural abnormalities.
Adult TAZKD mouse muscle shows abnormal mitochondrial structures with signs of mitophagy in addition to paracrystalline and honeycomb-like structures that are associated with altered lipid and phospholipid metabolism (1). In embryos, abnormal mitochondrial ultrastructure includes disrupted sarcomeres, vacuolated cristae, reduction in total mitochondrial area density, reduction in mitochondrial size, and lower cristae density (58). Ventricular and atrial muscle of TAZKD hearts show evidence of significant mitochondrial deterioration and mitophagy (1). Mitochondria also have large central vacuoles and abnormalities in mitochondria-associated membrane size and morphology (1). The mouse exhibits increased mtDNA content in heart and skeletal muscle, in accordance with mitochondrial proliferation (1).
Potential Therapies
See Table 4.
Table 4.
Therapeutic interventions
Human Barth Patients | TAZKD Mouse Model | |
---|---|---|
Gene replacement | ModRNA (full-length human tafazzin cDNA) in Barth iPSC-derived cardiomyocytes AAV-mediated gene therapy tested in dermal fibroblasts |
Cardiomyopathy reversible when DOX inducer is removed Des-TAZ vector gene therapy AAV9-TAZ gene therapy CL-Nanodisk as lipid replacement therapy |
Pharmacological | Bezafibrate tested in lymphocytes and current clinical trials Resveratrol tested in lymphocytes Elamipretide clinical trials MitoTempo (antioxidant) tested in iPSC-derived cardiomyocytes Linoleic acid tested in fibroblasts and iPSC-derived cardiomyocytes |
Bezafibrate ROS scavengers |
Cardiovascular | Angiotensin converting enzyme inhibitors (ACEi), angiotensin receptor blockers (ARB), β-blockers Berlin EXCOR device Heart transplant Implantable cardioverter defibrillators |
|
Skeletal myopathy | Exercise training | Endurance training |
Neutropenia | Granulocyte colony stimulating factor (G-CSF) | |
References | 13, 36, 65, 84, 85, 88, 92, 96, 99, 101 | 1, 32, 33, 71, 78, 88, 89, 92, 99 |
AAV, adeno-associated virus; CL, cardiolipin; DOX, doxycycline; iPSC, induced pluripotent stem cells; TAZ, tafazzin; TAZKD, knock down expression of tafazzin mRNA.
Based on the clinical and experimental data, potential therapeutic targets for Barth syndrome have included TAZ deficiency, altered CL profiles, and ROS, among others. As such, strategies such as gene therapy, CL protection, and antioxidants have been implemented in both the mouse model and clinical trials (Fig. 2). Unsurprisingly, the mouse “trials” have focused on the cardiac and muscular manifestations of the disease.
Fig. 2.
Gene replacement or gene editing therapy may be possible in the future, particularly with the refinement of CRISPR/cas9 approaches. Enzyme replacement therapy is another possible therapeutic approach. *Tested in TAZKD mouse model. †Tested in human Barth cells or in human clinical trials. AAV, adeno-associated virus; HF, heart failure; ND-CL, nanodisk-cardiolipin; OXPHOS, oxidative phosphorylation; PL, phospholipids; ROS, reactive oxygen species; TAZ, tafazzin.
Gene or lipid replacement therapy.
Conceptually, replacing the deficient enzyme of the abnormal lipid would “cure” Barth syndrome. Tafazzin is a nuclear-encoded enzyme that remodels CL to its mature, tetralinoleoyl form; abnormal tafazzin, therefore, leads to abnormal CL profiles. In this X-linked (Xq28) disorder, many different tafazzin splice variants and mutations have now been described. Pathogenic variants include missense, nonsense, splicing defects, full or partial deletions, and frame shifts (45). The human tafazzin gene produces four major mRNA splice variants, with two known to be functional: TAZ lacking exon 5 (Δ5) and full-length TAZ (45). Further evidence has suggested that the TAZ-Δ5 isoform is likely to be predominant and the most functional (49, 94, 104). Indeed, no disease-causing variant has been attributed to exon 5 to date (Barth Syndrome Foundation Human Tafazzin Gene Variants Database, accessed August 9, 2019: https://www.barthsyndrome.org/research/tafazzindatabase.html). A small percentage of patients may exhibit a milder clinical phenotype, correlating with milder biochemical abnormalities (11). However, no genotype-phenotype correlation has been demonstrated to date (38).
Indeed, withdrawal of doxycycline in the inducible TAZKD mouse can reverse tafazzin mRNA knockdown at least partially within 4 wk (1). Proof-of-concept studies of human Barth cell lines show that gene replacement therapy has great potential to mitigate or cure the disease.
Adeno-associated virus (AAV)-mediated gene replacement therapy has been proposed as a mechanism to compensate for the native dysfunctional tafazzin gene and restore mitochondrial and cardioskeletal function (88). The AAV vectors package the genetic promoter and gene of interest to provide stable gene transfer. Studies in the TAZKD model using AAV9-TAZ gene therapy have shown promise for clinical translation in the future (89). Suzuki-Hatano and coworkers (88, 89) reported both phenotypic rescue and normalization of proteomics profiles in TAZKD mice treated with AAV-mediated gene replacement. Phenotypic and biochemical rescue included several parameters of activity and exercise, cardiac function, mitochondrial ultrastructural changes, and mitochondrial functioning (88); adults seemed to respond very similarly to neonates. However, it should be noted that TAZKD mice were injected with AAV vectors as neonates and as 3-mo-old adult mice, which would be before the manifestations of any disease phenotype. Therefore, we believe what this group showed was prevention (not necessarily treatment) of Barth syndrome symptoms, biochemical abnormalities, and proteomic abnormalities in the TAZKD mouse model.
Nanodisk (ND) delivery particles encasing exogenous CL have also been attempted in the TAZKD mouse, based on in vitro cell lines studies (33). However, the 10-wk administration of ND-CL failed to induce any improvement in the MLCL-to-CL ratio. Therefore, AAV-mediated gene replacement to restore normal tafazzin (and therefore CL) appears to be a more viable approach.
Pharmacological therapies targeting consequences of gene mutation.
Peroxisome proliferator-activated receptors (PPARs) are members of the nuclear hormone receptor superfamily, act as ligand-activated transcription factors, play a central role in energy metabolism and mitochondrial bioenergetics, and have been studied as a target for Taz associated cardiac dysfunction. Bezafibrate is a pan-PPAR agonist (41). The rationale to study bezafibrate in Barth syndrome stemmed from 1) known effects on mitochondrial biogenesis (4); 2) effects on fatty acid metabolism in a disease in which fatty acid metabolism is dysregulated (12, 70); 3) a clinical trial of bezafibrate in human patients with a different mitochondrial disorder, CPT2 deficiency (10); and 4) salient effects on human Barth lymphocytes, including an improvement/amelioration of the MLCL-to-CL ratio (101). Treatment with the pan-PPAR agonist bezafibrate prevents development of LV dilation and LV systolic dysfunction in TAZKD mice (32, 71), and the PPAR pan-agonist bezafibrate ameliorated the cardiomyopathy in TAZKD mice, while increasing mitochondrial biogenesis (32). It has also been suggested that voluntary exercise potentiates the bezafibrate action in skeletal muscle (71), and that endurance training provides beneficial effects, such as increased mitochondrial content, decreased ROS production, and restoration of complex II activity in cardiac muscle (78). Surprisingly, the improvement in systolic function in TAZKD mice was accompanied by a decrease in CL in the heart, a finding attributed to the increased number of mitochondria with reduced content of normal CL. Nevertheless, it remains unclear why in vivo there was no improvement in the MLCL-to-CL ratio as had been seen in Barth lymphocytes; we speculate there may have been an improvement in β-oxidation and fatty acid metabolism.
ROS are another favored target for potential therapies. Therapies targeting ROS have been investigated in the TAZKD mouse by targeting catalase expression to mitochondria to relieve oxidative stress, by crossing the so-called MCAT mouse with the TAZKD mouse (37). Catalase reduces levels of H2O2 by converting this to water and oxygen. While this genetic approach improved oxidative stress, there were no significant effects on the development of cardiomyopathy in the TAZKD mouse. The results suggested a need for further investigation of ROS, both as a pathogenetic pathway and as a potential therapeutic target.
DISCUSSION
Altered profiles in the signature mitochondrial phospholipid CL have now been implicated in heart failure, including pediatric heart failure. Therefore, insights into the pathogenesis of Barth syndrome, which is the only disease in which CL is primarily affected, may offer insights into heart failure. Although many assume that limitations of energy (ATP) or bioenergetics arise from mitochondrial disorders such as Barth syndrome, especially in a highly metabolic organ such as the heart, the TAZKD mouse model has shed insights into mitochondrial functioning and pathogenesis of this rare disease that go beyond simple ATP.
Modern biomedical research relies on animal models of human disease to study disease pathogenesis and insights into potential targets for therapeutic intervention. For the past several years, the doxycycline-inducible, shRNA-mediated, tafazzin-knockdown mouse has represented the best mammalian model of human Barth syndrome. It is apparent from the review of the published data above that this mouse exhibits many of the physiological, cellular, and biochemical features of the human Barth syndrome (1, 42, 58, 59, 61, 79) and offers us some insights into several aspects of pediatric heart failure and cardiomyopathy, including prenatal-lethal cardiomyopathies. This mouse model has demonstrated, unsurprisingly, the importance of mitochondria and mitochondrial phospholipids in cardiac function, but there are also suggestions that normal mitochondrial functioning during development is essential for normal myocardial patterning, formation, and functioning. Altered bioenergetics does not seem to explain the entire picture. Nevertheless, the TAZKD model has not recapitulated the human disease perfectly. Moreover, biochemical and metabolic data have not been particularly insightful into the pathogenesis of the Barth cardiomyopathy, specifically, why the abnormal biochemistry, omics, and mitochondria lead to DCM, HCM, and/or LVNC.
Alterations in lipid metabolism are now well-known to coexist with, and contribute to, the pathogenesis of cardiac and skeletal muscle disorders (69). During heart failure, cardiac CL content is reduced (76) and the CL pool is altered such that a progressive decrease of mitochondrial function and energy metabolism is observed (22). Deficient CL biosynthesis and remodeling has been examined in both pediatric and adult heart failure patients and has demonstrated evidence of different mechanisms of disease between children and adults (14). Thus using an animal model with a similarly altered CL profile has the potential to lead to new therapeutic approaches for heart failure in children that differ from those used in adults.
Limitations of TAZKD Model
This is a knockdown—not a knockout—mouse model. Therefore, a small percentage of functional tafazzin is still present. Additionally, the inducible knockdown model requires an inducer, which may have confounding effects on the tissues. However, with the TAZKD model, the controls have also been fed doxycycline to account for potential confounders. Doxycycline itself may also be problematic, particularly in mitochondrial diseases, since mitochondria are bacteria that, eons ago, made their way into the cell in a symbiotic relationship. Lastly, the details of the animal model must be taken into account when analyzing results: in this case, there must be adequate time allowed for the knockdown of mRNA, protein, and finally CL, all of which precede disease manifestations or progression.
Future Directions
ROS is postulated by many to be an important contributing factor in the etiology and progression of many disease processes, including Barth syndrome (see above), but this view has been challenged in part by the failure of antioxidant therapy to consistently yield beneficial effects. The MCAT × TAZKD mouse crosses experiments outlined above provide one such example. Also, superoxide and H2O2 exert deleterious effects inside a cell by different mechanisms (35), so this may explain inconsistent results in “antioxidant” trials. The results indicate a need for further study in ROS and antioxidant therapy in relation to pediatric cardiomyopathies and heart failure. Downstream of metabolic changes, we really do not understand the link to cardiomyopathy and skeletal myopathy and other clinical features of Barth syndrome. Mouse models can be used to further analyze downstream consequences (and, therefore, potential therapeutic targets) of altered metabolism, including ROS.
Data based on the TAZKD model have shown that therapeutics using AAV vectors may be useful in the future for the subset of those pediatric cardiomyopathies with heart failure that are genetic in origin. It is interesting that trials have utilized the full-length human tafazzin cDNA, when it seems as if the Δ5 tafazzin isoform would be perhaps more appropriate. Similar considerations for the packaging of specific protein isoforms may need to be taken into account when applying AAV vector therapy to other pediatric genetic cardiomyopathies.
A new TAZKO mouse model has now been developed, and data are beginning to emerge on the pathogenesis of Barth syndrome, with an accumulation of unpublished data on its representativeness of the human condition. As such, an appraisal of the TAZKD mouse model will provide further context for the utility of this new model going forward. The movement toward a knockout model to represent the human condition of Barth syndrome could present differences in phenotype compared with the knockdown due to possible compensatory pathways in the knockout model not present in the TAZKD mouse (23). The inducible TAZKD model also has its own advantages that allow for time-dependent modeling of different diseases and studying of reversibility, neither of which is possible with a knockout. Especially when considering gene therapy, which will occur late in the disease, it has to be proven that the damage caused by the disease is reversible; therefore, requiring a knockdown model over a knockout. Thus the findings from the TAZKD mouse phenotype are significant in our understanding of altered CL profiles due to tafazzin deficiency and potential therapeutic approaches.
Conclusions
Despite some variations from the clinical human syndrome, the TAZKD mouse model of Barth syndrome has led to important insights into disease pathogenesis, especially the Barth cardiomyopathy. Given heart failure in children is most often due to cardiomyopathies (51), new insights into pediatric heart failure can be gleaned from our understanding of the TAZKD mammalian model that exhibits a primary cardiomyopathy. Such insights may bring us closer to identifying further therapeutic targets, including CL manipulation, which could translate to therapies for pediatric heart failure.
GRANTS
This study was supported by grants from the Barth Syndrome Foundation (to C. K. L. Phoon), a Dean’s Undergraduate Research Fund (to P. C. Miller), and the National Institutes of Health (to M. Schlame).
DISCLOSURES
M. Ren and M. Schlame are on the Scientific and Medical Advisory Board of the Barth Syndrome Foundation.
AUTHOR CONTRIBUTIONS
M.R. and C.K.P. conceived and designed research; M.R., P.C.M., M.S., and C.K.P. analyzed data; C.K.P. interpreted results of experiments; C.K.P. prepared figures; M.R., P.C.M., M.S., and C.K.P. drafted manuscript; M.R., P.C.M., M.S., and C.K.P. edited and revised manuscript; M.R., P.C.M., M.S., and C.K.P. approved final version of manuscript.
ACKNOWLEDGMENTS
Some of these data were presented at a poster session at the Barth Syndrome 9th International Scientific, Medical & Family Conference—“Power Up!” in Clearwater, FL, on July 21, 2018.
REFERENCES
- 1.Acehan D, Vaz F, Houtkooper RH, James J, Moore V, Tokunaga C, Kulik W, Wansapura J, Toth MJ, Strauss A, Khuchua Z. Cardiac and skeletal muscle defects in a mouse model of human Barth syndrome. J Biol Chem 286: 899–908, 2011. doi: 10.1074/jbc.M110.171439. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Acehan D, Xu Y, Stokes DL, Schlame M. Comparison of lymphoblast mitochondria from normal subjects and patients with Barth syndrome using electron microscopic tomography. Lab Invest 87: 40–48, 2007. doi: 10.1038/labinvest.3700480. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Angelini R, Lobasso S, Gorgoglione R, Bowron A, Steward CG, Corcelli A. Cardiolipin fingerprinting of leukocytes by MALDI-TOF/MS as a screening tool for Barth syndrome. J Lipid Res 56: 1787–1794, 2015. doi: 10.1194/jlr.D059824. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Augustyniak J, Lenart J, Gaj P, Kolanowska M, Jazdzewski K, Stepien PP, Buzanska L. Bezafibrate upregulates mitochondrial biogenesis and influence neural differentiation of human-induced pluripotent stem cells. Mol Neurobiol 56: 4346–4363, 2019. doi: 10.1007/s12035-018-1368-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Barth PG, Scholte HR, Berden JA, Van der Klei-Van Moorsel JM, Luyt-Houwen IE, Van ’t Veer-Korthof ET, Van der Harten JJ, Sobotka-Plojhar MA. An X-linked mitochondrial disease affecting cardiac muscle, skeletal muscle and neutrophil leucocytes. J Neurol Sci 62: 327–355, 1983. doi: 10.1016/0022-510X(83)90209-5. [DOI] [PubMed] [Google Scholar]
- 6.Barth PG, Van den Bogert C, Bolhuis PA, Scholte HR, van Gennip AH, Schutgens RB, Ketel AG. X-linked cardioskeletal myopathy and neutropenia (Barth syndrome): respiratory-chain abnormalities in cultured fibroblasts. J Inherit Metab Dis 19: 157–160, 1996. doi: 10.1007/BF01799418. [DOI] [PubMed] [Google Scholar]
- 7.Barth PG, Wanders RJ, Vreken P. X-linked cardioskeletal myopathy and neutropenia (Barth syndrome)-MIM 302060. J Pediatr 135: 273–276, 1999. doi: 10.1016/S0022-3476(99)70118-6. [DOI] [PubMed] [Google Scholar]
- 8.Bashir A, Bohnert KL, Reeds DN, Peterson LR, Bittel AJ, de Las Fuentes L, Pacak CA, Byrne BJ, Cade WT. Impaired cardiac and skeletal muscle bioenergetics in children, adolescents, and young adults with Barth syndrome. Physiol Rep 5: e13130, 2017. doi: 10.14814/phy2.13130. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Bissler JJ, Tsoras M, Göring HH, Hug P, Chuck G, Tombragel E, McGraw C, Schlotman J, Ralston MA, Hug G. Infantile dilated X-linked cardiomyopathy, G4.5 mutations, altered lipids, and ultrastructural malformations of mitochondria in heart, liver, and skeletal muscle. Lab Invest 82: 335–344, 2002. doi: 10.1038/labinvest.3780427. [DOI] [PubMed] [Google Scholar]
- 10.Bonnefont JP, Bastin J, Laforêt P, Aubey F, Mogenet A, Romano S, Ricquier D, Gobin-Limballe S, Vassault A, Behin A, Eymard B, Bresson JL, Djouadi F. Long-term follow-up of bezafibrate treatment in patients with the myopathic form of carnitine palmitoyltransferase 2 deficiency. Clin Pharmacol Ther 88: 101–108, 2010. doi: 10.1038/clpt.2010.55. [DOI] [PubMed] [Google Scholar]
- 11.Bowron A, Honeychurch J, Williams M, Tsai-Goodman B, Clayton N, Jones L, Shortland GJ, Qureshi SA, Heales SJ, Steward CG. Barth syndrome without tetralinoleoyl cardiolipin deficiency: a possible ameliorated phenotype. J Inherit Metab Dis 38: 279–286, 2015. [Erratum in J Inherit MetabDis 39: 151, 2016.] doi: 10.1007/s10545-014-9747-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Cade WT, Bohnert KL, Peterson LR, Patterson BW, Bittel AJ, Okunade AL, de Las Fuentes L, Steger-May K, Bashir A, Schweitzer GG, Chacko SK, Wanders RJ, Pacak CA, Byrne BJ, Reeds DN. Blunted fat oxidation upon submaximal exercise is partially compensated by enhanced glucose metabolism in children, adolescents, and young adults with Barth syndrome. J Inherit Metab Dis 42: 480–493, 2019. doi: 10.1002/jimd.12094. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Cade WT, Reeds DN, Peterson LR, Bohnert KL, Tinius RA, Benni PB, Byrne BJ, Taylor CL. Endurance exercise training in young adults with barth syndrome: a pilot study. JIMD Rep 32: 15–24, 2016. doi: 10.1007/8904_2016_553. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Chatfield KC, Sparagna GC, Sucharov CC, Miyamoto SD, Grudis JE, Sobus RD, Hijmans J, Stauffer BL. Dysregulation of cardiolipin biosynthesis in pediatric heart failure. J Mol Cell Cardiol 74: 251–259, 2014. doi: 10.1016/j.yjmcc.2014.06.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Chatzispyrou IA, Guerrero-Castillo S, Held NM, Ruiter JPN, Denis SW, IJlst L, Wanders RJ, van Weeghel M, Ferdinandusse S, Vaz FM, Brandt U, Houtkooper RH. Barth syndrome cells display widespread remodeling of mitochondrial complexes without affecting metabolic flux distribution. Biochim Biophys Acta Mol Basis Dis 1864: 3650–3658, 2018. doi: 10.1016/j.bbadis.2018.08.041. [DOI] [PubMed] [Google Scholar]
- 16.Clarke SL, Bowron A, Gonzalez IL, Groves SJ, Newbury-Ecob R, Clayton N, Martin RP, Tsai-Goodman B, Garratt V, Ashworth M, Bowen VM, McCurdy KR, Damin MK, Spencer CT, Toth MJ, Kelley RI, Steward CG. Barth syndrome. Orphanet J Rare Dis 8: 23, 2013. doi: 10.1186/1750-1172-8-23. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Cole LK, Kim JH, Amoscato AA, Tyurina YY, Bayir H, Karimi B, Siddiqui TJ, Kagan VE, Hatch GM, Kauppinen TM. Aberrant cardiolipin metabolism is associated with cognitive deficiency and hippocampal alteration in tafazzin knockdown mice. Biochim Biophys Acta Mol Basis Dis 1864: 3353–3367, 2018. doi: 10.1016/j.bbadis.2018.07.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Cole LK, Mejia EM, Vandel M, Sparagna GC, Claypool SM, Dyck-Chan L, Klein J, Hatch GM. Impaired cardiolipin biosynthesis prevents hepatic steatosis and diet-induced obesity. Diabetes 65: 3289–3300, 2016. doi: 10.2337/db16-0114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Das BB. Current State of Pediatric Heart Failure. Children (Basel) 5: E88, 2018. doi: 10.3390/children5070088. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Dolinsky VW, Cole LK, Sparagna GC, Hatch GM. Cardiac mitochondrial energy metabolism in heart failure: role of cardiolipin and sirtuins. Biochim Biophys Acta 1861: 1544–1554, 2016. doi: 10.1016/j.bbalip.2016.03.008. [DOI] [PubMed] [Google Scholar]
- 21.Dudek J, Cheng IF, Chowdhury A, Wozny K, Balleininger M, Reinhold R, Grunau S, Callegari S, Toischer K, Wanders RJ, Hasenfuß G, Brügger B, Guan K, Rehling P. Cardiac-specific succinate dehydrogenase deficiency in Barth syndrome. EMBO Mol Med 8: 139–154, 2016. doi: 10.15252/emmm.201505644. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Dudek J, Hartmann M, Rehling P. The role of mitochondrial cardiolipin in heart function and its implication in cardiac disease. Biochim Biophys Acta Mol Basis Dis 1865: 810–821, 2019. doi: 10.1016/j.bbadis.2018.08.025. [DOI] [PubMed] [Google Scholar]
- 23.El-Brolosy MA, Stainier DY. Genetic compensation: a phenomenon in search of mechanisms. PLoS Genet 13: e1006780, 2017. doi: 10.1371/journal.pgen.1006780. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Figarella-Branger D, Pellissier JF, Scheiner C, Wernert F, Desnuelle C. Defects of the mitochondrial respiratory chain complexes in three pediatric cases with hypotonia and cardiac involvement. J Neurol Sci 108: 105–113, 1992. doi: 10.1016/0022-510X(92)90195-Q. [DOI] [PubMed] [Google Scholar]
- 25.Fillmore N, Lopaschuk GD. The link between pediatric heart failure and mitochondrial lipids. J Mol Cell Cardiol 76: 71–72, 2014. doi: 10.1016/j.yjmcc.2014.08.002. [DOI] [PubMed] [Google Scholar]
- 26.Finsterer J. Barth syndrome: mechanisms and management. Appl Clin Genet 12: 95–106, 2019. doi: 10.2147/TACG.S171481. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Gaspard GJ, McMaster CR. Cardiolipin metabolism and its causal role in the etiology of the inherited cardiomyopathy Barth syndrome. Chem Phys Lipids 193: 1–10, 2015. doi: 10.1016/j.chemphyslip.2015.09.005. [DOI] [PubMed] [Google Scholar]
- 28.Griffiths EJ. Mitochondria and heart disease. Adv Exp Med Biol 942: 249–267, 2012. doi: 10.1007/978-94-007-2869-1_11. [DOI] [PubMed] [Google Scholar]
- 29.Hornby B, McClellan R, Buckley L, Carson K, Gooding T, Vernon HJ. Functional exercise capacity, strength, balance and motion reaction time in Barth syndrome. Orphanet J Rare Dis 14: 37, 2019. doi: 10.1186/s13023-019-1006-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Houtkooper RH, Turkenburg M, Poll-The BT, Karall D, Pérez-Cerdá C, Morrone A, Malvagia S, Wanders RJ, Kulik W, Vaz FM. The enigmatic role of tafazzin in cardiolipin metabolism. Biochim Biophys Acta 1788: 2003–2014, 2009. doi: 10.1016/j.bbamem.2009.07.009. [DOI] [PubMed] [Google Scholar]
- 31.Huang Y, Powers C, Madala SK, Greis KD, Haffey WD, Towbin JA, Purevjav E, Javadov S, Strauss AW, Khuchua Z. Cardiac metabolic pathways affected in the mouse model of barth syndrome. PLoS One 10: e0128561, 2015. doi: 10.1371/journal.pone.0128561. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Huang Y, Powers C, Moore V, Schafer C, Ren M, Phoon CK, James JF, Glukhov AV, Javadov S, Vaz FM, Jefferies JL, Strauss AW, Khuchua Z. The PPAR pan-agonist bezafibrate ameliorates cardiomyopathy in a mouse model of Barth syndrome. Orphanet J Rare Dis 12: 49, 2017. doi: 10.1186/s13023-017-0605-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Ikon N, Hsu FF, Shearer J, Forte TM, Ryan RO. Evaluation of cardiolipin nanodisks as lipid replacement therapy for Barth syndrome. J Biomed Res 32: 107–112, 2018. doi: 10.7555/JBR.32.20170094. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Ikon N, Ryan RO. Cardiolipin and mitochondrial cristae organization. Biochim Biophys Acta Biomembr 1859: 1156–1163, 2017. doi: 10.1016/j.bbamem.2017.03.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Imlay JA. Cellular defenses against superoxide and hydrogen peroxide. Annu Rev Biochem 77: 755–776, 2008. doi: 10.1146/annurev.biochem.77.061606.161055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Jefferies JL. Barth syndrome. Am J Med Genet C Semin Med Genet 163: 198–205, 2013. doi: 10.1002/ajmg.c.31372. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Johnson JM, Ferrara PJ, Verkerke AR, Coleman CB, Wentzler EJ, Neufer PD, Kew KA, de Castro Brás LE, Funai K. Targeted overexpression of catalase to mitochondria does not prevent cardioskeletal myopathy in Barth syndrome. J Mol Cell Cardiol 121: 94–102, 2018. doi: 10.1016/j.yjmcc.2018.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Kang SL, Forsey J, Dudley D, Steward CG, Tsai-Goodman B. Clinical characteristics and outcomes of cardiomyopathy in Barth syndrome: the UK experience. Pediatr Cardiol 37: 167–176, 2016. doi: 10.1007/s00246-015-1260-z. [DOI] [PubMed] [Google Scholar]
- 39.Kelley RI. Quantification of 3-methylglutaconic acid in urine, plasma, and amniotic fluid by isotope-dilution gas chromatography/mass spectrometry. Clin Chim Acta 220: 157–164, 1993. doi: 10.1016/0009-8981(93)90044-5. [DOI] [PubMed] [Google Scholar]
- 40.Kelley RI, Cheatham JP, Clark BJ, Nigro MA, Powell BR, Sherwood GW, Sladky JT, Swisher WP. X-linked dilated cardiomyopathy with neutropenia, growth retardation, and 3-methylglutaconic aciduria. J Pediatr 119: 738–747, 1991. doi: 10.1016/S0022-3476(05)80289-6. [DOI] [PubMed] [Google Scholar]
- 41.Khuchua Z, Glukhov AI, Strauss AW, Javadov S. Elucidating the beneficial role of PPAR agonists in cardiac diseases. Int J Mol Sci 19: 3464, 2018. doi: 10.3390/ijms19113464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Kiebish MA, Yang K, Liu X, Mancuso DJ, Guan S, Zhao Z, Sims HF, Cerqua R, Cade WT, Han X, Gross RW. Dysfunctional cardiac mitochondrial bioenergetic, lipidomic, and signaling in a murine model of Barth syndrome. J Lipid Res 54: 1312–1325, 2013. doi: 10.1194/jlr.M034728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Kim J, Lee K, Fujioka H, Tandler B, Hoppel CL. Cardiac mitochondrial structure and function in tafazzin-knockdown mice. Mitochondrion 43: 53–62, 2018. doi: 10.1016/j.mito.2018.10.005. [DOI] [PubMed] [Google Scholar]
- 44.Kimura T, Kimura AK, Ren M, Berno B, Xu Y, Schlame M, Epand RM. Substantial decrease in plasmalogen in the heart associated with tafazzin deficiency. Biochemistry 57: 2162–2175, 2018. doi: 10.1021/acs.biochem.8b00042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Kirwin SM, Manolakos A, Barnett SS, Gonzalez IL. Tafazzin splice variants and mutations in Barth syndrome. Mol Genet Metab 111: 26–32, 2014. doi: 10.1016/j.ymgme.2013.11.006. [DOI] [PubMed] [Google Scholar]
- 46.Koop AM, Hagdorn QA, Bossers GP, van Leusden T, Gerding A, van Weeghel M, Vaz FM, Koonen DP, Silljé HH, Berger RM, Bartelds B. Right ventricular pressure overload alters cardiac lipid composition. Int J Cardiol 287: 96–105, 2019. doi: 10.1016/j.ijcard.2019.04.004. [DOI] [PubMed] [Google Scholar]
- 47.Kulik W, van Lenthe H, Stet FS, Houtkooper RH, Kemp H, Stone JE, Steward CG, Wanders RJ, Vaz FM. Bloodspot assay using HPLC-tandem mass spectrometry for detection of Barth syndrome. Clin Chem 54: 371–378, 2008. doi: 10.1373/clinchem.2007.095711. [DOI] [PubMed] [Google Scholar]
- 48.Landriscina C, Megli FM, Quagliariello E. Turnover of fatty acids in rat liver cardiolipin: comparison with other mitochondrial phospholipids. Lipids 11: 61–66, 1976. doi: 10.1007/BF02532585. [DOI] [PubMed] [Google Scholar]
- 49.Lu YW, Galbraith L, Herndon JD, Lu YL, Pras-Raves M, Vervaart M, Van Kampen A, Luyf A, Koehler CM, McCaffery JM, Gottlieb E, Vaz FM, Claypool SM. Defining functional classes of Barth syndrome mutation in humans. Hum Mol Genet 25: 1754–1770, 2016. doi: 10.1093/hmg/ddw046. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Malhotra A, Edelman-Novemsky I, Xu Y, Plesken H, Ma J, Schlame M, Ren M. Role of calcium-independent phospholipase A2 in the pathogenesis of Barth syndrome. Proc Natl Acad Sci USA 106: 2337–2341, 2009. doi: 10.1073/pnas.0811224106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Masarone D, Valente F, Rubino M, Vastarella R, Gravino R, Rea A, Russo MG, Pacileo G, Limongelli G. Pediatric heart failure: a practical guide to diagnosis and management. Pediatr Neonatol 58: 303–312, 2017. doi: 10.1016/j.pedneo.2017.01.001. [DOI] [PubMed] [Google Scholar]
- 52.Mazzocco MM, Henry AE, Kelly RI. Barth syndrome is associated with a cognitive phenotype. J Dev Behav Pediatr 28: 22–30, 2007. doi: 10.1097/01.DBP.0000257519.79803.90. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.McKenzie M, Lazarou M, Thorburn DR, Ryan MT. Mitochondrial respiratory chain supercomplexes are destabilized in Barth Syndrome patients. J Mol Biol 361: 462–469, 2006. doi: 10.1016/j.jmb.2006.06.057. [DOI] [PubMed] [Google Scholar]
- 54.Moullan N, Mouchiroud L, Wang X, Ryu D, Williams EG, Mottis A, Jovaisaite V, Frochaux MV, Quiros PM, Deplancke B, Houtkooper RH, Auwerx J. Tetracyclines disturb mitochondrial function across eukaryotic models: a call for caution in biomedical research. Cell Rep 10: 1681–1691, 2015. doi: 10.1016/j.celrep.2015.02.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Murphy E, Ardehali H, Balaban RS, DiLisa F, Dorn GW 2nd, Kitsis RN, Otsu K, Ping P, Rizzuto R, Sack MN, Wallace D, Youle RJ; American Heart Association Council on Basic Cardiovascular Sciences, Council on Clinical Cardiology, and Council on Functional Genomics and Translational Biology . Mitochondrial function, biology, and role in disease: a scientific statement from the American Heart Association. Circ Res 118: 1960–1991, 2016. doi: 10.1161/RES.0000000000000104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Neustein HB, Lurie PR, Dahms B, Takahashi M. An X-linked recessive cardiomyopathy with abnormal mitochondria. Pediatrics 64: 24–29, 1979. [PubMed] [Google Scholar]
- 57.Ong SB, Hall AR, Hausenloy DJ. Mitochondrial dynamics in cardiovascular health and disease. Antioxid Redox Signal 19: 400–414, 2013. doi: 10.1089/ars.2012.4777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Phoon CK, Acehan D, Schlame M, Stokes DL, Edelman-Novemsky I, Yu D, Xu Y, Viswanathan N, Ren M. Tafazzin knockdown in mice leads to a developmental cardiomyopathy with early diastolic dysfunction preceding myocardial noncompaction. J Am Heart Assoc 1: e000455, 2012. doi: 10.1161/JAHA.111.000455. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Powers C, Huang Y, Strauss A, Khuchua Z. Diminished exercise capacity and mitochondrial bc1 complex deficiency in tafazzin-knockdown mice. Front Physiol 4: 74, 2013. doi: 10.3389/fphys.2013.00074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Price JF. Congestive heart failure in children. Pediatr Rev 40: 60–70, 2019. doi: 10.1542/pir.2016-0168. [DOI] [PubMed] [Google Scholar]
- 61.Ren M, Phoon CK, Schlame M. Metabolism and function of mitochondrial cardiolipin. Prog Lipid Res 55: 1–16, 2014. doi: 10.1016/j.plipres.2014.04.001. [DOI] [PubMed] [Google Scholar]
- 62.Ren M, Xu Y, Erdjument-Bromage H, Donelian A, Phoon CKL, Terada N, Strathdee D, Neubert TA, Schlame M. Extramitochondrial cardiolipin suggests a novel function of mitochondria in spermatogenesis. J Cell Biol 218: 1491–1502, 2019. doi: 10.1083/jcb.201808131. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Reynolds S, Kreider CM, Meeley LE, Bendixen RM. Taste perception and sensory sensitivity: relationship to feeding problems in boys with Barth Syndrome. J Rare Disord 3: 1–9, 2015. [PMC free article] [PubMed] [Google Scholar]
- 64.Rigaud C, Lebre AS, Touraine R, Beaupain B, Ottolenghi C, Chabli A, Ansquer H, Ozsahin H, Di Filippo S, De Lonlay P, Borm B, Rivier F, Vaillant MC, Mathieu-Dramard M, Goldenberg A, Viot G, Charron P, Rio M, Bonnet D, Donadieu J. Natural history of Barth syndrome: a national cohort study of 22 patients. Orphanet J Rare Dis 8: 70, 2013. doi: 10.1186/1750-1172-8-70. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Roberts AE, Nixon C, Steward CG, Gauvreau K, Maisenbacher M, Fletcher M, Geva J, Byrne BJ, Spencer CT. The Barth Syndrome Registry: distinguishing disease characteristics and growth data from a longitudinal study. Am J Med Genet A 158A: 2726–2732, 2012. doi: 10.1002/ajmg.a.35609. [DOI] [PubMed] [Google Scholar]
- 66.Romer AJ, Rajagopal SK, Kameny RJ. Initial presentation and management of pediatric heart failure. Curr Opin Pediatr 30: 319–325, 2018. doi: 10.1097/MOP.0000000000000624. [DOI] [PubMed] [Google Scholar]
- 67.Rossano JW, Jang GY. Pediatric heart failure: current state and future possibilities. Korean Circ J 45: 1–8, 2015. doi: 10.4070/kcj.2015.45.1.1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Rossano JW, Shaddy RE. Heart failure in children: etiology and treatment. J Pediatr 165: 228–233, 2014. doi: 10.1016/j.jpeds.2014.04.055. [DOI] [PubMed] [Google Scholar]
- 69.Saini-Chohan HK, Mitchell RW, Vaz FM, Zelinski T, Hatch GM. Delineating the role of alterations in lipid metabolism to the pathogenesis of inherited skeletal and cardiac muscle disorders: thematic review series: genetics of human lipid diseases. J Lipid Res 53: 4–27, 2012. doi: 10.1194/jlr.R012120. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Sandlers Y, Mercier K, Pathmasiri W, Carlson J, McRitchie S, Sumner S, Vernon HJ. Metabolomics reveals new mechanisms for pathogenesis in Barth syndrome and introduces novel roles for cardiolipin in cellular function. PLoS One 11: e0151802, 2016. doi: 10.1371/journal.pone.0151802. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Schafer C, Moore V, Dasgupta N, Javadov S, James JF, Glukhov AI, Strauss AW, Khuchua Z. The Effects of PPAR stimulation on cardiac metabolic pathways in Barth syndrome mice. Front Pharmacol 9: 318, 2018. doi: 10.3389/fphar.2018.00318. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Schlame M, Greenberg ML. Biosynthesis, remodeling and turnover of mitochondrial cardiolipin. Biochim Biophys Acta Mol Cell Biol Lipids 1862: 3–7, 2017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Schlame M, Kelley RI, Feigenbaum A, Towbin JA, Heerdt PM, Schieble T, Wanders RJ, DiMauro S, Blanck TJ. Phospholipid abnormalities in children with Barth syndrome. J Am Coll Cardiol 42: 1994–1999, 2003. doi: 10.1016/j.jacc.2003.06.015. [DOI] [PubMed] [Google Scholar]
- 74.Seibler J, Kleinridders A, Küter-Luks B, Niehaves S, Brüning JC, Schwenk F. Reversible gene knockdown in mice using a tight, inducible shRNA expression system. Nucleic Acids Res 35: e54, 2007. doi: 10.1093/nar/gkm122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Seneviratne AK, Xu M, Henao JJA, Fajardo VA, Hao Z, Voisin V, Xu GW, Hurren R, Kim S, MacLean N, Wang X, Gronda M, Jeyaraju D, Jitkova Y, Ketela T, Mullokandov M, Sharon D, Thomas G, Chouinard-Watkins R, Hawley JR, Schafer C, Yau HL, Khuchua Z, Aman A, Al-Awar R, Gross A, Claypool SM, Bazinet RP, Lupien M, Chan S, De Carvalho DD, Minden MD, Bader GD, Stark KD, LeBlanc P, Schimmer AD. The mitochondrial transacylase, tafazzin, regulates for AML stemness by modulating intracellular levels of phospholipids. Cell Stem Cell 24: 621–636.e16, 2019. [Erratum in Cell Stem Cell 24: 1007, 2019.] doi: 10.1016/j.stem.2019.02.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Sheeran FL, Pepe S. Mitochondrial bioenergetics and dysfunction in failing heart. Adv Exp Med Biol 982: 65–80, 2017. doi: 10.1007/978-3-319-55330-6_4. [DOI] [PubMed] [Google Scholar]
- 77.Shen Z, Ye C, McCain K, Greenberg ML. The role of cardiolipin in cardiovascular health. BioMed Res Int 2015: 891707, 2015. doi: 10.1155/2015/891707. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Soustek MS, Baligand C, Falk DJ, Walter GA, Lewin AS, Byrne BJ. Endurance training ameliorates complex 3 deficiency in a mouse model of Barth syndrome. J Inherit Metab Dis 38: 915–922, 2015. doi: 10.1007/s10545-015-9834-8. [DOI] [PubMed] [Google Scholar]
- 79.Soustek MS, Falk DJ, Mah CS, Toth MJ, Schlame M, Lewin AS, Byrne BJ. Characterization of a transgenic short hairpin RNA-induced murine model of Tafazzin deficiency. Hum Gene Ther 22: 865–871, 2011. doi: 10.1089/hum.2010.199. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Sparagna GC, Chicco AJ, Murphy RC, Bristow MR, Johnson CA, Rees ML, Maxey ML, McCune SA, Moore RL. Loss of cardiac tetralinoleoyl cardiolipin in human and experimental heart failure. J Lipid Res 48: 1559–1570, 2007. doi: 10.1194/jlr.M600551-JLR200. [DOI] [PubMed] [Google Scholar]
- 81.Spencer CT, Bryant RM, Day J, Gonzalez IL, Colan SD, Thompson WR, Berthy J, Redfearn SP, Byrne BJ. Cardiac and clinical phenotype in Barth syndrome. Pediatrics 118: e337–e346, 2006. doi: 10.1542/peds.2005-2667. [DOI] [PubMed] [Google Scholar]
- 82.Spencer CT, Byrne BJ, Bryant RM, Margossian R, Maisenbacher M, Breitenger P, Benni PB, Redfearn S, Marcus E, Cade WT. Impaired cardiac reserve and severely diminished skeletal muscle O2 utilization mediate exercise intolerance in Barth syndrome. Am J Physiol Heart Circ Physiol 301: H2122–H2129, 2011. doi: 10.1152/ajpheart.00479.2010. [DOI] [PubMed] [Google Scholar]
- 83.Spencer CT, Byrne BJ, Gewitz MH, Wechsler SB, Kao AC, Gerstenfeld EP, Merliss AD, Carboni MP, Bryant RM. Ventricular arrhythmia in the X-linked cardiomyopathy Barth syndrome. Pediatr Cardiol 26: 632–637, 2005. doi: 10.1007/s00246-005-0873-z. [DOI] [PubMed] [Google Scholar]
- 84.Steward CG. Treatment of Barth Syndrome by CARDIOlipin MANipulation (CARDIOMAN): a randomised placebo controlled pilot trial conducted by the nationally commissioned Barth Syndrome Service (Online). Bristol Trials Centre; http://cteu.bris.ac.uk/our-studies/?trialType=Rare-diseases [Accessed August 20, 2019]. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Steward CG, Groves SJ, Taylor CT, Maisenbacher MK, Versluys B, Newbury-Ecob RA, Ozsahin H, Damin MK, Bowen VM, McCurdy KR, Mackey MC, Bolyard AA, Dale DC. Neutropenia in Barth syndrome: characteristics, risks, and management. Curr Opin Hematol 26: 6–15, 2019. doi: 10.1097/MOH.0000000000000472. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Steward CG, Newbury-Ecob RA, Hastings R, Smithson SF, Tsai-Goodman B, Quarrell OW, Kulik W, Wanders R, Pennock M, Williams M, Cresswell JL, Gonzalez IL, Brennan P. Barth syndrome: an X-linked cause of fetal cardiomyopathy and stillbirth. Prenat Diagn 30: 970–976, 2010. doi: 10.1002/pd.2599. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Storch EA, Keeley M, Merlo LJ, St. Amant JB, Jacob M, Storch JF, Spencer C, Byrne BJ. Psychosocial functioning in youth with Barth syndrome. Child Health Care 38: 137–156, 2009. doi: 10.1080/02739610902813344. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Suzuki-Hatano S, Saha M, Rizzo SA, Witko RL, Gosiker BJ, Ramanathan M, Soustek MS, Jones MD, Kang PB, Byrne BJ, Cade WT, Pacak CA. AAV-mediated TAZ gene replacement restores mitochondrial and cardioskeletal function in Barth syndrome. Hum Gene Ther 30: 139–154, 2019. doi: 10.1089/hum.2018.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Suzuki-Hatano S, Saha M, Soustek MS, Kang PB, Byrne BJ, Cade WT, Pacak CA. AAV9-TAZ gene replacement ameliorates cardiac tmt proteomic profiles in a mouse model of Barth syndrome. Mol Ther Methods Clin Dev 13: 167–179, 2019. doi: 10.1016/j.omtm.2019.01.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Szczepanek K, Allegood J, Aluri H, Hu Y, Chen Q, Lesnefsky EJ. Acquired deficiency of tafazzin in the adult heart: impact on mitochondrial function and response to cardiac injury. Biochim Biophys Acta 1861: 294–300, 2016. doi: 10.1016/j.bbalip.2015.12.004. [DOI] [PubMed] [Google Scholar]
- 91.Towbin JA, Jefferies JL. Cardiomyopathies due to left ventricular noncompaction, mitochondrial and storage diseases, and inborn errors of metabolism. Circ Res 121: 838–854, 2017. doi: 10.1161/CIRCRESAHA.117.310987. [DOI] [PubMed] [Google Scholar]
- 92.Valianpour F, Wanders RJ, Overmars H, Vaz FM, Barth PG, van Gennip AH. Linoleic acid supplementation of Barth syndrome fibroblasts restores cardiolipin levels: implications for treatment. J Lipid Res 44: 560–566, 2003. doi: 10.1194/jlr.M200217-JLR200. [DOI] [PubMed] [Google Scholar]
- 93.van Werkhoven MA, Thorburn DR, Gedeon AK, Pitt JJ. Monolysocardiolipin in cultured fibroblasts is a sensitive and specific marker for Barth Syndrome. J Lipid Res 47: 2346–2351, 2006. doi: 10.1194/jlr.D600024-JLR200. [DOI] [PubMed] [Google Scholar]
- 94.Vaz FM, Houtkooper RH, Valianpour F, Barth PG, Wanders RJ. Only one splice variant of the human TAZ gene encodes a functional protein with a role in cardiolipin metabolism. J Biol Chem 278: 43089–43094, 2003. doi: 10.1074/jbc.M305956200. [DOI] [PubMed] [Google Scholar]
- 95.Vernon HJ, Sandlers Y, McClellan R, Kelley RI. Clinical laboratory studies in Barth Syndrome. Mol Genet Metab 112: 143–147, 2014. doi: 10.1016/j.ymgme.2014.03.007. [DOI] [PubMed] [Google Scholar]
- 96.Vernon HJ, Thompson WR, Hornby B. Elamipretide (ELAM) in subjects with Barth syndrome (TAZPOWER): a phase 2 randomized, double-blind, placebo-controlled crossover trial followed by an open-label treatment extension (Abstract). Circulation 138, Suppl 1: A16368, 2018. doi: 10.1161/circ.138.suppl_1.16368. [DOI] [Google Scholar]
- 97.Vreken P, Valianpour F, Nijtmans LG, Grivell LA, Plecko B, Wanders RJ, Barth PG. Defective remodeling of cardiolipin and phosphatidylglycerol in Barth syndrome. Biochem Biophys Res Commun 279: 378–382, 2000. doi: 10.1006/bbrc.2000.3952. [DOI] [PubMed] [Google Scholar]
- 98.Wahjudi PN, Yee JK, Martinez SR, Zhang J, Teitell M, Nikolaenko L, Swerdloff R, Wang C, Lee WN. Turnover of nonessential fatty acids in cardiolipin from the rat heart. J Lipid Res 52: 2226–2233, 2011. doi: 10.1194/jlr.M015966. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Wang G, McCain ML, Yang L, He A, Pasqualini FS, Agarwal A, Yuan H, Jiang D, Zhang D, Zangi L, Geva J, Roberts AE, Ma Q, Ding J, Chen J, Wang DZ, Li K, Wang J, Wanders RJ, Kulik W, Vaz FM, Laflamme MA, Murry CE, Chien KR, Kelley RI, Church GM, Parker KK, Pu WT. Modeling the mitochondrial cardiomyopathy of Barth syndrome with induced pluripotent stem cell and heart-on-chip technologies. Nat Med 20: 616–623, 2014. doi: 10.1038/nm.3545. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Woulfe KC, Bruns DR. From pediatrics to geriatrics: mechanisms of heart failure across the life-course. J Mol Cell Cardiol 126: 70–76, 2019. doi: 10.1016/j.yjmcc.2018.11.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Xu Y, Phoon CK, Berno B, D’Souza K, Hoedt E, Zhang G, Neubert TA, Epand RM, Ren M, Schlame M. Loss of protein association causes cardiolipin degradation in Barth syndrome. Nat Chem Biol 12: 641–647, 2016. doi: 10.1038/nchembio.2113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Xu Y, Schlame M. The turnover of glycerol and acyl moieties of cardiolipin. Chem Phys Lipids 179: 17–24, 2014. doi: 10.1016/j.chemphyslip.2013.10.005. [DOI] [PubMed] [Google Scholar]
- 103.Xu Y, Sutachan JJ, Plesken H, Kelley RI, Schlame M. Characterization of lymphoblast mitochondria from patients with Barth syndrome. Lab Invest 85: 823–830, 2005. [Erratum in Lab Invest 85: 831, 2005.] doi: 10.1038/labinvest.3700274. [DOI] [PubMed] [Google Scholar]
- 104.Xu Y, Zhang S, Malhotra A, Edelman-Novemsky I, Ma J, Kruppa A, Cernicica C, Blais S, Neubert TA, Ren M, Schlame M. Characterization of tafazzin splice variants from humans and fruit flies. J Biol Chem 284: 29230–29239, 2009. doi: 10.1074/jbc.M109.016642. [DOI] [PMC free article] [PubMed] [Google Scholar]