Next Article in Journal
Enhancement of Antioxidant Systems and Storability of Two Plum Cultivars by Preharvest Treatments with Salicylates
Next Article in Special Issue
Event-Associated Oxygen Consumption Rate Increases ca. Five-Fold When Interictal Activity Transforms into Seizure-Like Events In Vitro
Previous Article in Journal
Determination of Highly Sensitive Biological Cell Model Systems to Screen BPA-Related Health Hazards Using Pathway Studio
Previous Article in Special Issue
Understanding the Epilepsy in POLG Related Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mitochondrial Liver Toxicity of Valproic Acid and Its Acid Derivatives Is Related to Inhibition of α-Lipoamide Dehydrogenase

1
Department of Epileptology and Life & Brain Center, University of Bonn, Sigmund-Freud-Str. 25, D-53105 Bonn, Germany
2
Institute for Drug Research, School of Pharmacy, Faculty of Medicine, The Hebrew University of Jerusalem, Jerusalem 91120, Israel
3
Institute for Research in Military Medicine, Faculty of Medicine, The Hebrew University of Jerusalem, Jerusalem 91120, Israel
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2017, 18(9), 1912; https://doi.org/10.3390/ijms18091912
Submission received: 28 July 2017 / Revised: 1 September 2017 / Accepted: 3 September 2017 / Published: 6 September 2017

Abstract

:
The liver toxicity of valproic acid (VPA) is an established side effect of this widely used antiepileptic drug, which is extremely problematic for patients with metabolic epilepsy and particularly epilepsy due to mitochondrial dysfunction. In the present report, we investigated the reason for liver mitochondrial toxicity of VPA and several acid and amide VPA analogues. While the pyruvate and 2-oxoglutarate oxidation rates of rat brain mitochondria were nearly unaffected by VPA, rat liver mitochondrial pyruvate and 2-oxoglutarate oxidation was severely impaired by VPA concentrations above 100 µM. Among the reactions involved in pyruvate oxidation, pyruvate transport and dehydrogenation steps were not affected by VPA, while α-lipoamide dehydrogenase was strongly inhibited. Strong inhibition of α-lipoamide dehydrogenase was also noted for the VPA one-carbon homolog sec-butylpropylacetic acid (SPA) and to a lesser extent for the VPA constitutional isomer valnoctic acid (VCA), while the corresponding amides of the above three acids valpromide (VPD), sec-butylpropylacetamide (SPD) and valnoctamide (VCD) showed only small effects. We conclude that the active inhibitors of pyruvate and 2-oxoglutarate oxidation are the CoA conjugates of VPA and its acid analogues affecting selectively α-lipoamide dehydrogenase in liver. Amide analogues of VPA, like VCD, show low inhibitory effects on mitochondrial oxidative phosphorylation in the liver, which might be relevant for treatment of patients with mitochondrial epilepsy.

Graphical Abstract

1. Introduction

Valproic acid (VPA; Figure 1) is a widely used classical antiepileptic drug with a high responder rate especially in patients with genetic generalized epilepsies (GGE, former nomenclature—idiopathic generalized epilepsies). One well known side effect of VPA is its pronounced liver toxicity, which is relevant in genetic epilepsies affecting the brain and also the liver. One typical example of an epileptic disorder when the liver toxicity of VPA is extremely relevant is Alpers-Hüttenlocher syndrome caused by mutation in mitochondrial DNA-polymerase γ [1,2]. At present, the detailed reason for VPA’s strong liver toxicity is not very well established, but a strong inhibition of pyruvate oxidation of liver mitochondria by VPA has been noted [3,4]. Current concepts propose sequestration of mitochondrial coenzyme A (CoA) [5,6,7], inhibition of pyruvate transport [8], inhibition of ATP- and GTP-dependent succinate:CoA ligases [9], inhibition of hepatic N-acetylglutamate synthase [10], and inhibition of α-lipoamide dehydrogenase [11].
To avoid the side effects including liver toxicity and teratogenicity and to retain its wide antiepileptic spectrum of efficacy of the drug, several derivatives of VPA have been developed. The amide of the chiral constitutional isomer of valpromide (VPD; Figure 1; the amide of VPA) valnoctamide (VCD; Figure 1) is a central nervous system (CNS)-active drug [12,13,14] that shows a low biotransformation to the corresponding acid-valnoctic acid (VCA) [12,15,16]. This is different from VPD, which undergoes rapid transformation to VPA in humans. From 1964 until 2005 in several European countries, VCD (racemate) was commercially available as the anxiolytic drug Nirvanil® [13,17]. In its chemical structure, VCD has two stereogenic centers (Figure 1). The mixture of both stereoisomers of VCD (racemate) has been reported to have a wide spectrum of anticonvulsant activity at concentrations that are 2–16 times lower than VPA and are dependent on the epilepsy model [13]. When intraperitoneally administered at seizure onset, 65 mg/kg VCD protected against pilocarpine-induced status epilepticus (SE) in the rat model [18]. However, this protection was lost when administered at 80 mg/kg 30 min after seizure onset [19], but it did block the pilocarpine-induced electrographic SE at the higher dose of 180 mg/kg [20,21]. In contrast to VPA, racemic-VCD, its corresponding acid VCA, and two of its individual stereoisomers, (2R,3S)-VCD and (2S,3S)-VCD, did not show any significant teratogenic effects in SWV/Finn mice—an inbred mouse strain that is highly susceptible to VPA-induced teratogenicity [22,23]. sec-Butylpropylacetamide (SPD, Figure 1) is a one-carbon homologue of valnoctamide (VCD, Figure 1). SPD has been recently reported by us to show a unique and broad-spectrum antiepileptic profile, which is better (lower ED50 values) than VPA and even superior to VCD in some epilepsy models [19,20]. Additionally, SPD blocked the behavioral and electrographic SE induced by the muscarinic agonist pilocarpine, and the organophosphates soman and paraoxon and showed in vivo neuroprotection that was associated with cognitive sparing [19,20,24,25].
In the present work, we investigated in detail the toxic effects of VPA and its corresponding CNS-active analogues on different pathways of oxidative phosphorylation of isolated mitochondria from rat liver and brain. From the obtained IC50 values we conclude that the inhibitory effects of CoA conjugates of VPA and its acid analogues are due to inhibition of liver α-lipoamide dehydrogenase.

2. Results

2.1. The Pyruvate Oxidation of Rat Liver Mitochondria Is Selectively Inhibited by Valproic Acid (VPA) and Its Acid Derivatives

First, we repeated the observations of previous investigators [3] showing the strong and selective inhibition of ADP-stimulated pyruvate oxidation of isolated rat liver mitochondria by valproic acid with an apparent IC50 value of about 50 μM (Figure 2B, red diamonds). In order to achieve stable and reproducible inhibition, we preincubated the mitochondria for 3 min in the presence of the drug and of 1 mM ATP. Interestingly, the same preincubation procedure resulted in almost no inhibition of ADP-stimulated pyruvate oxidation for rat brain mitochondria in a comparable concentration range of VPA (Figure 2A, red diamonds). On the other hand, the amide of valproic acid (VPD) was much less effective in inhibiting the pyruvate oxidation of rat liver mitochondria under similar experimental conditions (Figure 2B, blue circles), having an apparent IC50 value of 5 mM. It has been convincingly shown that VPA can be activated by the intramitochondrial medium chain acyl-CoA synthetase to valproyl-CoA, which is then very slowly metabolized by branched chain acyl-CoA dehydrogenase (Acadsb), followed by β-oxidation [26]. Therefore, our experiments support the conclusion that the intramitochondrially formed CoA-conjugate of VPA is essential for toxicity, since amide derivatives of VPA under our experimental conditions cannot form potentially toxic acyl CoA-derivatives. In brain mitochondria, the presence of CoA transferases (which are absent in liver) and the very low expression of intramitochondrial medium chain acyl-CoA synthetase keep the toxic valproyl-CoA concentrations at low levels.
In additional experiments, we tested the effects of the other VPA derivatives and their corresponding amides—SPA (sec-butylpropylacetic acid) and SPD; VCA and VCD—on their potential to inhibit pyruvate oxidation of rat mitochondria. The acid derivatives were effective, but less potent than VPA—SPA had an IC50 of 50 µM, and VCA had an IC50 of 900 µM (Figure 3A,B, red diamonds). The corresponding amides (Figure 3A,B, blue circles) were again much less effective in inhibiting pyruvate oxidation.
In further experiments, we tested the potential toxic effects of VPA and its derivatives on ADP-stimulated oxidation rates of octanoyl carnitine, glutamate, and succinate in isolated rat liver mitochondria. As presented in the Figures S1–S3 (Supplementary Materials), we observed only slight inhibitory effects at concentrations of VPA and its derivatives in the range below 1 mM (below 40% inhibition of respiration). That excludes significant toxic effects not directly related to the pyruvate oxidation pathway, like the proposed sequestration of CoA conjugates [5,6,7].

2.2. Pyruvate Dehydrogenase and Pyruvate Transport of Rat Liver Mitochondria Are Not Affected by VPA

In a next series of experiments, we studied which particular reactions involved in pyruvate oxidation might be specifically inhibited by VPA-CoA. The first possibility, already suggested by Turnbull et al. [27], is the dehydrogenation reaction of pyruvate. To evaluate the potential effect of VPA-CoA on this reaction, we preincubated mitochondria with VPA under activating conditions (1 mM ATP present) and then assessed the pyruvate dehydrogenation velocity in the presence of excess α-lipoamide dehydrogenase. As shown in Figure 4A, preincubation of mitochondria with 10 mM VPA and 1 mM ATP resulted even in a slight increase of this particular reaction.
In further experiments, we studied pyruvate transport, which has also been suggested to be responsible for VPA toxicity [8]. To do this, we studied the pyruvate-dependent swelling of mitochondria in ammonium-ion containing medium. This reaction was clearly inhibited by 1 mM cinnamate, a selective inhibitor of pyruvate transport (Figure 4B, blue and black traces), verifying specificity. Pretreatment of mitochondria with 1 mM VPA and 1 mM ATP, however, had no effect on the swelling velocity (Figure 4B, pink trace), indicating the absence of effects on pyruvate transport.

2.3. Valproic Acid and Their Acid Derivatives Inhibit α-Lipoamide Dehydrogenase

In a further set of experiments, we investigated the potential effects of VPA and its derivatives on α-lipoamide dehydrogenase. To do this with isolated mitochondria, we determined the pyruvate-dependent re-reduction of NAD(P)+ after the addition of oxidized α-lipoamide (Figure 5A).
In the absence of valproic acid, there is a substantial re-reduction of NAD(P)+ (black trace), which is strongly inhibited by 0.5 and 1 mM VPA, respectively (red and blue traces). This effect is highly specific for pyruvate (or 2-oxoglutarate) as substrate, since NAD(P)+ re-reduction by glutamate after the addition of oxidized α-lipoamide is not observed, and no effects of VPA are detected (Figure 5B). With this assay, we analyzed the inhibitory effects of all VPA derivatives tested in the prior respiration assays with pyruvate as substrate (Figure 6A–C). From the dose dependencies, it can be clearly seen that the sensitivity profile of α-lipoamide dehydrogenase reaction to the VPA derivatives is very similar to the sensitivity profile of pyruvate-dependent oxygen consumption (Table 1, Figure 2B and Figure 3A,B).
Since α-lipoamide dehydrogenase is also involved in the 2-oxoglutate dehydrogenase complex, we tested the sensitivity of ADP-stimulated 2-oxoglutarate oxidation rates of rat liver mitochondria to VPA and its analogues. As shown in the Figure S4A–C (Supplementary Materials), we observed an apparent IC50 value of 50 µM under comparable preincubation conditions for VPA and for the acid analogues of VPA—VCA an apparent IC50 of 200 µM and for SPA an apparent IC50 of 50 µM. These values and the observed lower inhibition effects of the corresponding amides were very similar to the values observed for pyruvate oxidation (Table 1, Figure 2B and Figure 3A,B). Therefore, it can be concluded that the strong inhibitory effects of VPA and its analogues on pyruvate and 2-oxoglutarate oxidation of rat liver mitochondria can be solely explained by inhibition of α-lipoamide dehydrogenase.

3. Discussion

The liver toxicity of VPA is one of the most relevant adverse side effects of this broad-spectrum antiepileptic drug. It has been noticed from clinical observations that the liver toxicity of VPA is extremely problematic in Alpers-Hüttenlocher syndrome—a mitochondrial form of epilepsy due to mutations in the mitochondrial DNA polymerase γ affecting brain and liver [1,2,28,29]. However, the detailed reason for the particular exacerbation of this adverse side effect of VPA in certain forms of mitochondrial epilepsy is not clear. To explain this problem, we conducted a detailed study of the in vitro toxicity of VPA and of CNS-active VPA analogues in isolated rat liver mitochondria. Our data clearly show that the pyruvate and 2-oxoglutarate oxidation rates of liver mitochondria are very effectively inhibited by VPA and its acid analogues SPA and VCA (VPA has an IC50 values of about 50 µM, SPA of 50 µM, and VCA of 200–900 µM), if the compounds are preincubated in the presence of ATP, allowing the intramitochondrial formation of the corresponding acyl CoA intermediate [4]. Similar to VPA, both SPA and VCA are also very likely activated by intramitochondrial medium chain acyl-CoA synthetase and slowly metabolized by branched chain acyl-CoA dehydrogenase, followed by β-oxidation [26].
Our experimental data are consistent with a selective inhibition of liver α-lipoamide dehydrogenase by valproyl-CoA or the CoA esters of its acid analogues due to the following findings: (i) Pyruvate dehydrogenase and pyruvate transport are almost not affected by preincubation of rat liver mitochondria with VPA in the mM range; (ii) ADP-stimulated octanoylcarnitine and glutamate oxidation of rat liver mitochondria is only slightly affected by preincubation with VPA in the µM range, which excludes considerable CoA sequestration effects; (iii) NAD(P)+ re-reduction from α-lipoamide and pyruvate is strongly inhibited by preincubation with VPA and its acid analogues in the µM range; and (iv) both the ADP-stimulated pyruvate and 2-oxoglutarate oxidation rates of rat liver mitochondria are strongly inhibited by preincubation with VPA and its acid analogues in the µM range. The liver specificity of the observed effects—the preincubation of rat brain mitochondria with VPA in the mM range had almost no effect on pyruvate (Figure 2A) and 2-oxoglutarate oxidation rates—appears to be related to the missing activity of CoA transferases in liver mitochondria, which precludes an efficient turnover of slowly metabolizing acyl CoAs. Moreover, intramitochondrial medium chain acyl-CoA synthetases allowing the ATP-dependent formation of CoA conjugates of VPA and its acid analogues have in brain very low expression levels (cf.: http://www.gtexportal.org/home/gene). The presented experimental data cannot exclude potential inhibitory effects of β-oxidation intermediates of VPA and its acid analogues on liver α-lipoamide dehydrogenase. Their concentrations reach only very low levels in comparison to valproyl-CoA [4], which makes this alternative explanation less likely.
The therapeutic blood levels of VPA in patients with epilepsy are in the range of 40–100 mg/L, which corresponds to a concentration range of 277–693 µM [30]. Therefore, the very high sensitivity of liver α-lipoamide dehydrogenase with an apparent IC50 of 80µM to VPA could convincingly explain why, under the condition of mitochondrial DNA depletion in the liver of Alpers-Hüttenlocher patients, the already limited capacity of mitochondrial oxidative phosphorylation would be further impaired by strong inhibition of this enzyme affecting both pyruvate dehydrogenase and 2-oxoglutarate dehydrogenase complexes. Other potential reasons for mitochondrial toxicity of VPA, like CoA sequestration CoA [5,6,7], the inhibition of pyruvate transport [8], the inhibition of ATP- and GTP-dependent succinate:CoA ligases [9], and the inhibition of hepatic N-acetylglutamate synthase [10], seem to not be relevant, since their apparent IC50 values are all the mM range.
It has to be mentioned that our study is mostly relevant for valproate-associated adverse effects in liver reported for mitochondrial forms of epilepsy. Under these circumstances, liver mitochondria with genetically impaired oxidative phosphorylation [28] are additionally challenged by inhibition of α-lipoamide dehydrogenase, which obviously leads to an amplification of toxic effects. It cannot be excluded that, in other forms of epilepsy, there are still additional relevant targets of VPA-associated adverse effects.
Among the tested acid analogues of VPA, VCA showed the lowest mitochondrial toxicity in liver. The corresponding amides—VPD, SPD, and VCD—that are not suitable for a fast, direct conversion to the corresponding CoA intermediate under our in vitro conditions, are much less potent mitochondrial toxins. In this context, it has to be noted that VPD undergoes a rapid biotransformation to VPA under in vivo conditions in humans, while VCD appears to be much more resistant to amidases [16]. Unlike SPD, which is currently in a preclinical stage of development, VCD recently underwent a phase IIb clinical trial in bipolar disorder, but the interim analysis was not successful [24,31]. Nevertheless, due to its metabolic stability (i.e., minimal conversion to its corresponding acid) [16], safety in humans, low liver mitochondrial toxicity (this report), and lack of teratogenicity [24,32], VCD might be a good alternative for VPA in patients with epilepsy, particularly for women of child bearing age and in patients under the risk of severe liver involvement. However, this still remains to be shown in further in vivo experimental studies in animals and in a successful clinical trial in patients.

4. Materials and Methods

4.1. Materials

VPA and all major chemicals were commercially obtained from Sigma-Aldrich (St. Louis, MO, USA). VCD, VCA, SPD and SPA were synthesized using previously described methods [19,21].

4.2. Isolation of Rat Brain and Rat Liver Mitochondria

We isolated mitochondria from a single rat brain using the protocol of Rosenthal et al. [33], which was slightly modified to obtain better characteristics [34]. Rat liver mitochondria were prepared according to Steinbrecht and Kunz [35] with the following modifications. The liver of one 50–60 days old Wistar rat (approx. 5 g) was immediately transferred into ice-cold solution A (0.3 M sucrose, 3 mM EGTA, pH 7.4) and shaken to wash out blood. Then we minced the liver, added around 30 mL of solution A, and homogenized it at 600 units/s using a potter homogenizer (Potter S, B. Braun Melsungen, Melsungen, Germany). Thereafter, the homogenate was centrifuged at 900× g for 5 min. The supernatant was passed through a cheesecloth and centrifuged at 12,000× g for 10 min. The resulting pellet was dissolved with approx. 20 mL of ice-cold solution B (0.3 sucrose, pH 7.4 adjusted with small amounts of Tris-base). The solution was transferred to a small glass homogenizer and homogenized 10–12 times manually. Finally, the suspension was centrifuged at 12,000× g for 10 min, and the resulting pellet was dissolved in the proportion 160 µL solution B per 1 g liver wet weight.

4.3. Determination of Substrate Oxidation Rates

Substrate oxidation rates were determined using high resolution respirometry. Rat brain and liver mitochondria (0.04–0.05 mg protein/mL for brain or 0.2 mg/mL protein/mL for liver) were incubated for 3 min in air gassed medium (10 mM KH2PO4, 60 mM KCl, 60 mM Tris-HCl, 110 mM mannitol, and 0.5 mM EDTA (pH 7.4)) in the presence of 5 mM MgCl2, 1 mM ATP and the respective amount of VPA or its analogues at 30 °C. Than the maximal substrate-supported oxygen uptake was measured by high-resolution respirometry using an Oroboros oxygraph [36] in the presence of different substrates and 2 mM ADP.

4.4. Determination of Pyruvate Dehydrogenase (PDH) Activity and Measurement of Mitochondrial Pyruvate Uptake

To determine the pyruvate dehydrogenase (PDH) activity and pyruvate transport, rat liver mitochondria (2.8–3.6 mg protein/mL) were incubated in the Oxygraph chamber in MTP medium in the presence of 5 mM MgCl2, 1 mM ATP, and the respective amount of VPA or its analogues for 3 min. PDH activity in this suspension was determined according to Scislowski and Davis [37] with the following modifications: the mitochondrial suspension (0.3 mg protein) was placed into a 0.5 mL spectrophotometric cuvette, containing 0.1 M potassium phosphate buffer (pH 7.4), 5 mM MgCl2, 3 mM NAD+, 0.1 mM CoASH, 2 mM dithiothreitol, 0.1% N-dodecyl-β-d-maltoside, 0.4 mM thiamine pyrophosphate, 20 µg/mL diaphorase (Boehringer Mannheim, Mannheim, Germany), and 10 mM pyruvate. PDH activity was measured at 30 °C using a spectrophotometer (Aminco DW 2000, SLM Instruments, Urbana, IL, USA) with 340–380 nm dual wavelength photometry.
For pyruvate transport determinations, the mitochondrial suspension (0.6–0.7 mg protein) was placed into a 3 mL spectrophotometric cuvette, containing 5 mM MgCl2, 2 µM rotenone, 1.5 µg/mL antimycin, and 0.1 M ammonium pyruvate (obtained from ammonium hydroxide and pyruvic acid). When indicated, 1 mM cinnamate was added [38]. Mitochondrial swelling was estimated at 30 °C using a spectrophotometer (Aminco DW 2000, SLM Instruments) at 546 nm.

4.5. Determination of α-Lipoamide Dehydrogenase Activity in Intact Mitochondria

α-Lipoamide dehydrogenase activity of rat liver mitochondria was analyzed, probing the reversed reaction of the enzyme [39]. After 3 min incubation of rat liver mitochondria (0.3 mg protein/mL) in MTP medium in the presence of 5 mM MgCl2, 1 mM ATP, and the respective amounts of VPA (or its derivatives) in 1 mL fluorimetric cuvette, the following additions were subsequently made: 5 mM malate and 10 mM pyruvate, 6.7 µM rotenone, 4 µM KCN, 75 µM α-lipoamide (30 mM stock solution was dissolved in 50% alcohol), and 10 mM arsenite, 0.4 µM 4,5,6,7-tetrachloro-2-trifluoromethyl benzimidazole (TTFB). The NAD(P)H reduction was monitored at λex = 340 nm, λem = 450 nm in a Shimadzu RF-5001PC (Shimadzu Scientific, Columbia, MD, USA) spectrofluorophotometer. DLDH activity was estimated from the NAD(P)H fluorescence increase after the addition of α-lipoamide. In agreement with Luis et al. [11], complete inhibition of DLDH activity was observed after the application of 10 mM arsenite.

Supplementary Materials

Supplementary materials can be found at www.mdpi.com/1422-0067/18/9/1912/s1.

Acknowledgments

This work was supported by the Deutsche Forschungsgemeinschaft (KU 911/21-1 and KU 911/21-2 to Wolfram S. Kunz, approval date: 4 February 2014 and 2 August 2017) and the Stiftung für Medizinische Wissenschaft (Frankfurt am Main).

Author Contributions

Wolfram S. Kunz and Christian E. Elger conceived and designed the experiments; Alexei P. Kudin performed the experiments; Alexei P. Kudin and Wolfram S. Kunz analyzed the data; Hafiz Mawasi, Arik Eisenkraft, and Meir Bialer contributed materials and reagents; Wolfram S. Kunz wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

VPAValproic acid
VPDValpromide
VCAValnoctic acid
VCDValnoctamide
SPAsec-Butylpropylacetic acid
SPDsec-Butylpropylacetamide
CNSCentral nervous system
PDHPyruvate dehydrogenase
SEStatus epilepticus

References

  1. Stewart, J.D.; Horvath, R.; Baruffini, E.; Ferrero, I.; Bulst, S.; Watkins, P.B.; Fontana, R.J.; Day, C.P.; Chinnery, P.F. Polymerase γ gene POLG determines the risk of sodium valproate-induced liver toxicity. Hepatology 2010, 52, 1791–1796. [Google Scholar] [CrossRef] [PubMed]
  2. Saneto, R.P.; Lee, I.C.; Koenig, M.K.; Bao, X.; Weng, S.W.; Naviaux, R.K.; Wong, L.J. POLG DNA testing as an emerging standard of care before instituting valproic acid therapy for pediatric seizure disorders. Seizure 2010, 19, 140–146. [Google Scholar] [CrossRef] [PubMed]
  3. Silva, M.F.; Ruiter, J.P.; Illst, L.; Jakobs, C.; Duran, M.; de Almeida, I.T.; Wanders, R.J. Valproate inhibits the mitochondrial pyruvate-driven oxidative phosphorylation in vitro. J. Inherit. Metab. Dis. 1997, 20, 397–400. [Google Scholar] [CrossRef] [PubMed]
  4. Silva, M.F.; Ruiter, J.P.; IJlst, L.; Allers, P.; ten Brink, H.J.; Jakobs, C.; Duran, M.; de Almeida, I.T.; Wanders, R.J. Synthesis and intramitochondrial levels of valproyl-coenzyme A metabolites. Anal. Biochem. 2001, 290, 60–67. [Google Scholar] [CrossRef] [PubMed]
  5. Becker, C.M.; Harris, R.A. Influence of valproic acid on hepatic carbohydrate and lipid metabolism. Arch. Biochem. Biophys. 1983, 223, 381–392. [Google Scholar] [CrossRef]
  6. Swartzentruber, M.S.; Harris, R.A. Inhibition of metabolic processes by coenzyme-A-sequestering aromatic acids. Prevention by para-chloro- and para-nitrobenzoic acids. Biochem. Pharmacol. 1987, 36, 3147–3153. [Google Scholar] [CrossRef]
  7. Ponchaut, S.; van Hoof, F.; Veitch, K. In vitro effects of valproate and valproate metabolites on mitochondrial oxidations: Relevance of CoA sequestration to the observed inhibitions. Biochem. Pharmacol. 1992, 43, 2435–2442. [Google Scholar] [CrossRef]
  8. Aires, C.C.; Soveral, G.; Luís, P.B.; ten Brink, H.J.; de Almeida, I.T.; Duran, M.; Wanders, R.J.; Silva, M.F. Pyruvate uptake is inhibited by valproic acid and metabolites in mitochondrial membranes. FEBS Lett. 2008, 582, 3359–3366. [Google Scholar] [CrossRef] [PubMed]
  9. Luís, P.B.; Ruiter, J.; IJlst, L.; de Almeida, I.T.; Duran, M.; Wanders, R.J.; Silva, M.F. Valproyl-CoA inhibits the activity of ATP- and GTP-dependent succinate:CoA ligases. J. Inherit. Metab. Dis. 2014, 37, 353–357. [Google Scholar] [CrossRef] [PubMed]
  10. Aires, C.C.; van Cruchten, A.; Ijlst, L.; de Almeida, I.T.; Duran, M.; Wanders, R.J.; Silva, M.F. New insights on the mechanisms of valproate-induced hyperammonemianhibition of hepatic N-acetylglutamate synthase activity by valproyl-CoA. J. Hepatol. 2011, 55, 426–434. [Google Scholar] [CrossRef] [PubMed]
  11. Luís, P.B.; Ruiter, J.P.; Aires, C.C.; Soveral, G.; de Almeida, I.T.; Duran, M.; Wanders, R.J.; Silva, M.F. Valproic acid metabolites inhibit dihydrolipoyl dehydrogenase activity leading to impaired 2-oxoglutarate-driven oxidative phosphorylation. Biochim. Biophys. Acta 2007, 1767, 1126–1133. [Google Scholar] [CrossRef] [PubMed]
  12. Barel, S.; Yagen, B.; Schurig, V.; Soback, S.; Pisani, F.; Perucca, E.; Bialer, M. Stereoselective pharmacokinetic analysis of valnoctamide in healthy subjects and in patients with epilepsy. Clin. Pharmacol. Ther. 1997, 61, 442–449. [Google Scholar] [CrossRef]
  13. Bialer, M.; Yagen, B. Valproic acid—Second generation. Neurotherapeutics 2007, 4, 130–137. [Google Scholar] [CrossRef] [PubMed]
  14. Bialer, M. Chemical properties of antiepileptic drugs (AEDs). Adv. Drug Deliv. Rev. 2007, 64, 887–895. [Google Scholar] [CrossRef] [PubMed]
  15. Bialer, M.; Haj-Yehia, A.; Barzaghi, N.; Pisani, F.; Perucca, E. Pharmacokinetics of a valpromide isomer, valnoctamide, in healthy subjects. Eur. J. Clin. Pharmacol. 1990, 38, 289–291. [Google Scholar] [CrossRef] [PubMed]
  16. Bialer, M. Clinical pharmacology of valpromide. Clin. Pharmacokinet. 1991, 20, 114–122. [Google Scholar] [CrossRef] [PubMed]
  17. Bialer, M.; White, H.S. Key factors in the discovery and development of new antiepileptic drugs (AEDs). Nat. Rev. Drug Discov. 2010, 9, 68–83. [Google Scholar] [CrossRef] [PubMed]
  18. Curia, G.; Longo, D.; Biagini, G.; Jones, R.S.; Avoli, M. The pilocarpine model of temporal lobe epilepsy. J. Neurosci. Methods 2008, 172, 143–157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. White, H.S.; Alex, A.B.; Pollock, A.; Hen, N.; Shekh-Ahmad, T.; Wilcox, K.S.; McDonough, J.H.; Stables, J.P.; Kaufmann, D.; Yagen, B.; et al. A new derivative of valproic acid amide possesses a broad-spectrum antiseizure profile and unique activity against status epilepticus and organophosphate neuronal damage. Epilepsia 2012, 53, 134–146. [Google Scholar] [CrossRef] [PubMed]
  20. Pouliot, W.; Bialer, M.; Hen, N.; Shekh-Ahmad, T.; Kaufmann, D.; Yagen, B.; Ricks, K.; Roach, B.; Nelson, C.; Dudek, F.E. Electrographic analysis of the effect of sec-butyl-propylacetamide on pharmacoresistant status epilepticus. Neuroscience 2013, 231, 145–156. [Google Scholar] [CrossRef] [PubMed]
  21. Kaufmann, D.; Yagen, B.; Minert, A.; Wlodarczyk, B.; Finnell, R.H.; Schurig, V.; Devor, M.; Bialer, M. Evaluation of the antiallodynic, teratogenic and pharmacokinetic profile of stereoisomers of valnoctamide, an amide derivative of a chiral isomer of valproic acid. Neuropharmacology 2010, 52, 1228–1236. [Google Scholar] [CrossRef] [PubMed]
  22. Shekh-Ahmad, T.; Hen, N.; Yagen, B.; McDonough, J.H.; Finnell, R.H.; Wlodarczyk, B.J.; Bialer, M. Stereoselective anticonvuslant and pharmacokinetic analysis of valnoctamide, a CNS-active derivative of valproic acid with low teratogenic potential. Epilepsia 2014, 55, 353–361. [Google Scholar] [CrossRef] [PubMed]
  23. Finnell, R.H.; Bennett, G.D.; Karras, S.B.; Mohl, V.K. Common hierarchies of susceptibility to the induction of neural tube defects in mouse embryos by valproic acid and its 4-propyl-4-pentenoic acid metabolite. Teratology 1988, 38, 313–320. [Google Scholar] [CrossRef] [PubMed]
  24. Bialer, M.; Johannessen, S.I.; Levy, R.H.; Perucca, E.; Tomson, T.; White, H.S. Progress report on new antiepileptic drugs: A summary of the Eleventh Eilat Conference (EILAT XI). Epilepsy Res. 2013, 103, 2–30. [Google Scholar] [CrossRef] [PubMed]
  25. Bar-Klein, G.; Swissa, E.; Kamintsky, L.; Shekh-Ahmad, T.; Saar-Ashkenazy, R.; Hubary, Y.; Shrot, S.; Stetlander, L.; Eisenkraft, A.; Friedman, A.; et al. sec-Butyl-propylacetamide (SPD) and two of its stereoisomers rapidly terminate paraoxon-induced status epilepticus in rats. Epilepsia 2014, 55, 1953–1958. [Google Scholar] [CrossRef] [PubMed]
  26. Luís, P.B.; Ruiter, J.P.; Ijlst, L.; de Almeida, I.T.; Duran, M.; Mohsen, A.W.; Vockley, J.; Wanders, R.J.; Silva, M.F. Role of isovaleryl-CoA dehydrogenase and short branched-chain acyl-CoA dehydrogenase in the metabolism of valproic acidmplications for the branched-chain amino acid oxidation pathway. Drug Metab. Dispos. 2011, 39, 1155–1160. [Google Scholar] [CrossRef] [PubMed]
  27. Turnbull, D.M.; Bone, A.J.; Bartlett, K.; Koundakjian, P.P.; Sherratt, H.S. The effect of valproate on intermediary metabolism in isolated rat hepatocytes and intact rats. Biochem. Pharmacol. 1983, 32, 1887–1892. [Google Scholar] [CrossRef]
  28. Zsurka, G.; Baron, M.; Stewart, J.D.; Kornblum, C.; Bös, M.; Sassen, R.; Taylor, R.W.; Elger, C.E.; Chinnery, P.F.; Kunz, W.S. Clonally expanded mitochondrial DNA mutations in epileptic individuals with mutated DNA polymerase γ. J. Neuropathol. Exp. Neurol. 2008, 67, 857–866. [Google Scholar] [CrossRef] [PubMed]
  29. Zsurka, G.; Kunz, W.S. Mitochondrial dysfunction and seizures: The neuronal energy crisis. Lancet Neurol. 2015, 14, 956–966. [Google Scholar] [CrossRef]
  30. Levy, R.H.; Shen, D.D.; Abbott, F.S.; Riggs, K.W.; Hachad, H. Valproic acidchemistry, biotransformaton and pharmacokinetics. In Antiepileptic Drugs, 5th ed.; Levy, R.H., Mattson, R.H., Meldrum, B.S., Perucca, W., Eds.; Lippincott, Williams & Wilkins: Philadelphia, PA, USA, 2002; pp. 780–800. [Google Scholar]
  31. Bersudsky, Y.; Applebaum, J.; Gaiduk, Y.; Sharony, L.; Mishory, A.; Podberezsky, A.; Agam, G.; Belmaker, R.H. Valnoctamide as a valproate substitute with low teratogenic potential in mania: A double-blind, controlled, add-on clinical trial. Bipolar Disord. 2010, 12, 376–382. [Google Scholar] [CrossRef] [PubMed]
  32. Wlodarczyk, B.J.; Ogle, K.; Lin, L.Y.; Bialer, M.; Finnell, R.H. Comparative teratogenicity analysis of valnoctamide, risperidone, and olanzapine in mice. Bipolar Disord. 2015, 17, 615–625. [Google Scholar] [CrossRef] [PubMed]
  33. Rosenthal, R.E.; Hamud, F.; Fiskum, G.; Varghese, P.J.; Sharpe, S. Cerebral ischemia and reperfusionrevention of brain mitochondrial injury by lidoflazine. J. Cereb. Blood Flow Metab. 1987, 7, 752–758. [Google Scholar] [CrossRef] [PubMed]
  34. Kudin, A.P.; Bimpong-Buta, N.Y.; Vielhaber, S.; Elger, C.E.; Kunz, W.S. Characterization of superoxide-producing sites in isolated brain mitochondria. J. Biol. Chem. 2004, 279, 4127–4135. [Google Scholar] [CrossRef] [PubMed]
  35. Steinbrecht, I.; Kunz, W. Use of “cycling” technic for random quantitative determination of the degree of reduction of NAD and NADP system in rat liver mitochondria with continuous recording of the measurements. Acta Biol. Med. Ger. 1970, 25, 731–747. [Google Scholar] [PubMed]
  36. Wisniewski, E.; Kunz, W.S.; Gellerich, F.N. Phosphate affects the distribution of flux control among the enzymes of oxidative phosphorylation in rat skeletal muscle mitochondria. J. Biol. Chem. 1993, 268, 9343–9346. [Google Scholar] [PubMed]
  37. Scislowski, P.W.D.; Davis, E.J. A sensitive spectrophotometric assay of pyruvate dehydrogenase activity. Anal. Biochem. 1986, 155, 400–404. [Google Scholar] [CrossRef]
  38. Eboli, M.L.; Paradies, G.; Galeotti, T.; Papa, S. Pyruvate transport in tumour-cell mitochondria. Biochim. Biophys. Acta 1977, 460, 183–187. [Google Scholar] [CrossRef]
  39. Patel, M.S.; Hong, Y.S. Lipoic acid as an antioxidant. The role of dihydrolipoamide dehydrogenase. Methods Mol. Biol. 1998, 108, 337–346. [Google Scholar] [PubMed]
Figure 1. Chemical structures of central nervous system (CNS)-active derivatives of valproic acid. The stars indicate the stereogenic centers of VCD and SPD.
Figure 1. Chemical structures of central nervous system (CNS)-active derivatives of valproic acid. The stars indicate the stereogenic centers of VCD and SPD.
Ijms 18 01912 g001
Figure 2. Inhibition of mitochondrial pyruvate oxidation by valproic acid in rat brain mitochondria (A) and rat liver mitochondria (B). Mitochondria (0.04–0.05 mg protein/mL for brain or 0.2 mg/mL protein/mL for liver) were preincubated for 3 min in presence of 1 mM ATP, 5 mM MgCl2 with the indicated amount of VPA. The maximal rate of respiration was determined in presence of 10 mM pyruvate, 5 mM malate, and 1 mM ADP. The plotted rates are averages of three independent experiments. VPD
Figure 2. Inhibition of mitochondrial pyruvate oxidation by valproic acid in rat brain mitochondria (A) and rat liver mitochondria (B). Mitochondria (0.04–0.05 mg protein/mL for brain or 0.2 mg/mL protein/mL for liver) were preincubated for 3 min in presence of 1 mM ATP, 5 mM MgCl2 with the indicated amount of VPA. The maximal rate of respiration was determined in presence of 10 mM pyruvate, 5 mM malate, and 1 mM ADP. The plotted rates are averages of three independent experiments. VPD
Ijms 18 01912 g002
Figure 3. Inhibition of pyruvate oxidation of isolated rat liver mitochondria by derivatives of valproic acid. (A) Red diamonds, SPA; blue circles, SPD. (B) Red diamonds, VCA; blue circles, VCD. Experimental conditions as indicated in the legend to Figure 2.
Figure 3. Inhibition of pyruvate oxidation of isolated rat liver mitochondria by derivatives of valproic acid. (A) Red diamonds, SPA; blue circles, SPD. (B) Red diamonds, VCA; blue circles, VCD. Experimental conditions as indicated in the legend to Figure 2.
Ijms 18 01912 g003
Figure 4. Effect of VPA on pyruvate dehydrogenase activity (A) and pyruvate transport activity (B) of isolated rat liver mitochondria. Liver mitochondria (2.8–3.6 mg protein/mL) were preincubated for 3 min in presence of 1 mM ATP, 5 mM MgCl2 with the indicated amount of VPA. (A) The plotted pyruvate dehydrogenase (PDH) activities are averages of three independent experiments; (B) Swelling of mitochondria in 0.1 M ammonium pyruvate buffer. One representative experiment (out of 3) is shown. Red trace: control; pink trace: 1 mM VPA; blue trace: 1 mM cinnamate; black trace: 1 mM VPA and 1 mM cinnamate. Experimental conditions as described in Materials and Methods.
Figure 4. Effect of VPA on pyruvate dehydrogenase activity (A) and pyruvate transport activity (B) of isolated rat liver mitochondria. Liver mitochondria (2.8–3.6 mg protein/mL) were preincubated for 3 min in presence of 1 mM ATP, 5 mM MgCl2 with the indicated amount of VPA. (A) The plotted pyruvate dehydrogenase (PDH) activities are averages of three independent experiments; (B) Swelling of mitochondria in 0.1 M ammonium pyruvate buffer. One representative experiment (out of 3) is shown. Red trace: control; pink trace: 1 mM VPA; blue trace: 1 mM cinnamate; black trace: 1 mM VPA and 1 mM cinnamate. Experimental conditions as described in Materials and Methods.
Ijms 18 01912 g004
Figure 5. Effect of VPA on pyruvate (A) and glutamate (B) dependent arsenite-sensitive re-reduction of NAD(P)+ by α-lipoamide dehydrogenase. Representative experimental traces of mitochondrial NAD(P)H fluorescence showing re-reduction of NAD(P)+ after α-lipoamide (75 µM) additions (first arrow). α-Lipoamide dehydrogenase is blocked by addition of 10 mM arsenite (second arrow). Finally, the uncoupler 4,5,6,7-tetrachloro-2-trifluoromethyl benzimidazole (TTFB, third arrow) was added. The detailed experimental conditions are described in the Materials and Methods section.
Figure 5. Effect of VPA on pyruvate (A) and glutamate (B) dependent arsenite-sensitive re-reduction of NAD(P)+ by α-lipoamide dehydrogenase. Representative experimental traces of mitochondrial NAD(P)H fluorescence showing re-reduction of NAD(P)+ after α-lipoamide (75 µM) additions (first arrow). α-Lipoamide dehydrogenase is blocked by addition of 10 mM arsenite (second arrow). Finally, the uncoupler 4,5,6,7-tetrachloro-2-trifluoromethyl benzimidazole (TTFB, third arrow) was added. The detailed experimental conditions are described in the Materials and Methods section.
Ijms 18 01912 g005
Figure 6. Dose-dependency of effects of VPA and its derivatives on pyruvate-dependent reduction of NAD(P)+ by α-lipoamide dehydrogenase. (A) Red diamonds, VPA; blue circles, VPD. (B) Red diamonds, SPA; blue circles, SPD. (C) Red diamonds, VCA; blue circles, VCD. Experimental conditions as indicated in Figure 5.
Figure 6. Dose-dependency of effects of VPA and its derivatives on pyruvate-dependent reduction of NAD(P)+ by α-lipoamide dehydrogenase. (A) Red diamonds, VPA; blue circles, VPD. (B) Red diamonds, SPA; blue circles, SPD. (C) Red diamonds, VCA; blue circles, VCD. Experimental conditions as indicated in Figure 5.
Ijms 18 01912 g006
Table 1. Apparent IC50 values of inhibition of different oxidation reactions of isolated rat liver mitochondria by VPA and its analogues.
Table 1. Apparent IC50 values of inhibition of different oxidation reactions of isolated rat liver mitochondria by VPA and its analogues.
VPAVPDVCAVCDSPASPD
pyruvate oxidation50 µM5 mM900 µM8 mM50 µM3 mM
2-oxoglutarate oxidation50 µM8 mM200 µM5 mM50 µM1 mM
glutamate oxidation10 mM3 mM2 mM4.5 mM2 mM2 mM
α-lipoamide dehydrogenase80 µM>1 mM 1300 µM>1 mM 1100 µM>1 mM 1
1 Estimated value. VPA: valproic acid; VPD: valpromide; VCA: valnoctic acid; VCD: valnoctamide; SPA: sec-butylpropylacetic acid; SPD: sec-butylpropylacetamide.

Share and Cite

MDPI and ACS Style

Kudin, A.P.; Mawasi, H.; Eisenkraft, A.; Elger, C.E.; Bialer, M.; Kunz, W.S. Mitochondrial Liver Toxicity of Valproic Acid and Its Acid Derivatives Is Related to Inhibition of α-Lipoamide Dehydrogenase. Int. J. Mol. Sci. 2017, 18, 1912. https://doi.org/10.3390/ijms18091912

AMA Style

Kudin AP, Mawasi H, Eisenkraft A, Elger CE, Bialer M, Kunz WS. Mitochondrial Liver Toxicity of Valproic Acid and Its Acid Derivatives Is Related to Inhibition of α-Lipoamide Dehydrogenase. International Journal of Molecular Sciences. 2017; 18(9):1912. https://doi.org/10.3390/ijms18091912

Chicago/Turabian Style

Kudin, Alexei P., Hafiz Mawasi, Arik Eisenkraft, Christian E. Elger, Meir Bialer, and Wolfram S. Kunz. 2017. "Mitochondrial Liver Toxicity of Valproic Acid and Its Acid Derivatives Is Related to Inhibition of α-Lipoamide Dehydrogenase" International Journal of Molecular Sciences 18, no. 9: 1912. https://doi.org/10.3390/ijms18091912

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop