Abstract
The control of fatty acid translocation across the mitochondrial membrane is mediated by the carnitine palmitoyltransferase (CPT) system. Modulation of its functionality has simultaneous effects on fatty acid and glucose metabolism. This encourages use of the CPT system as drug target for reduction of gluconeogenesis and restoration of lipid homeostasis, which are beneficial in the treatment of type 2 diabetes mellitus and obesity. Recently, crystal structures of CPT-2 were determined in uninhibited forms and in complexes with inhibitory substrate-analogs with anti-diabetic properties in animal models and in clinical studies. The CPT-2 crystal structures have advanced understanding of CPT structure–function relationships and will facilitate discovery of novel inhibitors by structure-based drug design. However, a number of unresolved questions regarding the biochemistry and pharmacology of CPT enzymes remain and are addressed in this review.
Keywords: Type 2 diabetes mellitus, Carnitine palmitoyltransferase, Enzyme isoforms, Malonyl-CoA, Drug discovery
Roles of CPTs in metabolic disease mechanisms
Type 2 diabetes mellitus (T2D) is a metabolic disorder characterized by resistance of peripheral target tissues, especially skeletal muscle, adipose tissue, and liver, to insulin and concomitant relative insulin deficiency. The occurrence of T2D is highly correlated with an increased risk of cardiovascular disease (atherosclerosis, diabetic retinopathy), neuropathy (sensory loss, diabetic foot disease), and nephropathy (kidney failure rendering patients dependent on hemodialysis). These complications emerge because of chronic hyperglycemia, which causes tissue damage by non-enzymatic glycation of proteins, lipids, and DNA [1]. The vastly growing incidence of T2D worldwide (diabetes epidemic) [2] and its severe health consequences have to be met by pharmacologic intervention.
Today several treatments are established for T2D that are applied depending on disease progression. Clinically mild presentations of T2D can be treated by established anti-diabetic drugs such as insulin secretagogs, exemplified by the more classic sulfonylurea compounds, the recently developed DPP-IV inhibitors, and insulin sensitizing thiazolidinediones (TZDs), which are agonists of the nuclear hormone receptor PPARγ, and metformin [3]. Advanced T2D with pancreatic failure requires administration of exogenous insulin. Several members of the TZD class of compounds have been approved for treatment of T2D. However, some TZDs have had to be withdrawn from clinical development and the use of approved substances had to be restricted because of hepatotoxicity and potential cardiotoxicity [3, 4]. In addition, it seems that weight gain is a common side effect of treatment with TZDs.
Strategies for the development of an improved treatment of T2D include the search for novel pathways and molecular targets involved in the control of glucose homeostasis. Innovative drug molecules should be highly efficacious and administered orally with simple dose regimen. As mono-therapy or in combination with established therapies, they should effectively ameliorate the symptoms of T2D and, therefore, delay or avoid the use of insulin treatment. One such novel strategy of T2D treatment—with no drug molecule currently on the market—is inhibition of the carnitine palmitoyltransferase (CPT) system (EC 2.3.1.21).
The CPT system consists of several components:
CPT-1, an integral transmembrane protein of the mitochondrial outer membrane that catalyzes the transfer of acyl moieties from co-enzyme A (CoA) to carnitine;
the carnitine–acylcarnitine translocase (CACT) transporter located in the mitochondrial inner membrane; and
CPT-2, a monotopic membrane protein on the matrix side of the mitochondrial inner membrane that reverses the reaction catalyzed by CPT-1.
These consecutive transesterifications enable the import of long-chain fatty acids (LCFA) into mitochondria (Fig. 1).
Fig. 1.
Model of acyl translocation by the CPT system. CPT-1 is located on the outer mitochondrial membrane and delivers acylcarnitine to CPT-2 via the transporter CACT, which are both located on the inner mitochondrial membrane
The acyl-CoA released by CPT-2 is channeled into β-oxidation and ultimately yields acetyl-CoA. In mitochondria of hepatocytes the acetyl-CoA derived from β-oxidation is a potent allosteric activator of pyruvate carboxylase, which catalyzes the first step of gluconeogenesis. In diabetic patients whose insulin-sensitive tissues are deprived of glucose, lipogenesis in the adipose tissue and β-oxidation in the liver occur at increased rates. The increased substrate flux through the CPT system together with prevailing glucagonergic effects in insulin-resistant T2D further suppress glycolysis and provide redox equivalents needed for gluconeogenesis.
To effectively treat T2D by exploiting the CPT system, two presumably synergistic approaches have been outlined:
inhibition of the CPT system in the liver with subsequent reduction of the gluconeogenesis rate; and
mobilization of ectopic lipid depots by stimulating the activity of the CPT system in skeletal muscle of obese patients, which would lead to increased insulin sensitivity [5].
For drug discovery the second approach is challenging because of the low feasibility of finding and optimizing compounds that activate the catalysis of an enzyme compared with active-site inhibitors, but progress in tools and methods for drug discovery might render such mechanisms tractable. The existence of organ-specific isoforms of CPT-1 expressed in liver (L-CPT-1) and muscle (M-CPT-1) might enable tackling of both mechanisms. A third isoform of CPT-1 expressed predominantly in brain tissue (B-CPT-1) was discovered in 2002 by Price et al. [6], but initially no catalytic activity could be detected in yeast extracts with heterologously expressed B-CPT-1. Later it was shown that B-CPT-1 KO mice show reduced food intake and body weight but have an increased susceptibility to becoming obese on a high-fat diet when compared with wild-type mice [7]. Recently, Sierra et al. [8] demonstrated that B-CPT-1 has, in fact, catalytic activity and is located in the endoplasmic reticulum. Because B-CPT-1 is co-expressed with L-CPT-1 in the retina [9], the latter is most likely the relevant mitochondrial isoform in brain neurons; this has also been shown by expression tests with northern blotting in rat hypothalamus [10]. In contrast, CPT-2 exists only as a single isoform that is ubiquitously expressed (Fig. 2).
Fig. 2.
Overview of CPT proteins and known isoforms. In contrast with the ubiquitous CPT-2 isoform, CPT-1 is differentially expressed in different tissues. More than one CPT-1 isoform may be present in one tissue (see text for details)
With the characterization of four different CPT isoforms and the identification of the CPT system as drug target, the next challenge is the discovery and development of molecules that can selectively modify the individual enzymatic activities of the CPT system. An important step was the isolation of emeriamine (i.e., aminocarnitine) derivatives, fungal metabolites that were shown to inhibit CPT-1 and have hypoglycemic activity [11, 12]. The development of a series of generic, aminocarnitine-based CPT inhibitors with anti-diabetic potential [13–16] provided proof of concept in respect of reduction of glucose levels in diabetic animal models. A recent study demonstrated that l-aminocarnitine can also inhibit CPT-2 in cultured cells [17]. On the basis of this result it has to be assumed that aminocarnitine derivatives can freely equilibrate over all components of the CPT system. Such unselective inhibition would clearly provoke undesired side-effects and, therefore, new chemical entities with improved specificity profiles need to be found. In this review we will focus on biochemical, structural, and pharmacological data that may contribute to the development of specific modulators of the various CPT isoforms.
Biochemical and biophysical data of CPT isoforms
CPT sources and reconstitution
The existence of separate CPT-1 and CPT-2 activities on either side of mitochondrial membranes was proposed as early as 1963 [18], but it took three decades to confirm this proposal by cloning and characterization of the individual enzymes [19, 20]. Considerable research effort was required to distinguish biochemically between CPT-1 and CPT-2 [21] and their (partial) purification and characterization as individual proteins [22–25].
CPT-2 can be purified by detergent extraction from native tissues and expressed with high yield in Escherichia coli [26]. In contrast, CPT-1 needs to be reconstituted in liposomes in order to recover pure and active enzyme from expression in Pichia pastoris [27]. Mitoplast preparations from the yeast Saccharomyces cerevisiae expressing CPT-1 also contain enzyme that has properties similar to that derived from native sources with regard to malonyl sensitivity and, importantly, its correct membrane topology [28, 29]. The yeast-expression systems have the advantage that they do not have any endogenous CPT activity, which enables investigation of mutant forms of CPT. In a recent review Zammit [30] summarized data demonstrating that in diabetic animal models the lipid composition and, as a result of changes in cholesterol contents, the fluidity of the mitochondrial outer membrane affects CPT-1 activity. This emphasizes the difficulty in characterizing the activity and inhibitor-sensitivity of CPT isoforms in vitro, because the choice of lipid or detergent for reconstitution (see below) may lock the enzyme in a state that is not representative of the metabolic condition of interest.
Quaternary structures
Comprehensive data on the degree of oligomerization of CPT enzymes in preparations from various native tissues and expression systems have been published. Interpretation of the results has been complicated by the need to use detergents. The enhanced quality of protein preparations and the refinement of preparative sizing technologies have improved the precision and accuracy of the investigation of CPT quaternary structures. A historic overview of the published data is given in Table 1.
Table 1.
Solution molecular weights of CPT-1 and CPT-2
CPT source | Extraction/solubilization detergent | Assay/detergent (v/v) | App. molecular weight [Da] | Ref. |
---|---|---|---|---|
CPT-1 | ||||
NA | NA | Gel filtration/0.15% Tween 20 | 150,000a | [25] |
Rat liver | 0.1% Tween 80 | Gel filtration/0.1% Tween 80 | 430,000 | [106] |
Recombinant rat L-CPT-1 from S. cerevisiae | 1% Triton X-100 or 1% Digitonin |
Gel filtration/1% Triton X-100 or native PAGE/ (1) 1% Triton X-100 (2) 2% Digitonin |
275,000 (trimer) or 550,000–620,000 (hexamer = dimer of trimers) | [40] |
Recombinant human L-CPT-1 (aa 168-778) from E. coli | 0.1% Fos12 | Analytical ultracentrifugation/0.1% Fos12 | 97% monomerb | Roche, unpublished results |
Recombinant human M-CPT-1 (aa 168–778) from E. coli | 0.1% DDM | Analytical ultracentrifugation/0.1% DDM | 70% monomer and 25% dimerb | Roche, unpublished results |
CPT-2 | ||||
Bovine liver | Endogenous lipids | Gel filtration/endogenous lipids | 68,000 | [24] |
NA | NA | Gel filtration/0.15% Tween 20 | 76,000a | [25] |
Rat liver | 0.1% Tween 80 | Gel filtration/0.1% Tween 80 | 430,000 | [106] |
Bovine heart | 6% Triton X-100/2% Triton X-100 | Gel filtration/1% Triton X-100 | 510,000 | [35] |
Rat liver | 0.5% Triton X-100 | Gel filtration/0.1% Triton X-100 or 0.5% Tween 20 | 280,000–300,000 | [107] |
Bovine liver | None | Gel filtration/none | 63,500 ± 800 | [108] |
Bovine heart | 6–0.002% Triton X-100, 25–12 mM βOG | Gel filtration, 12 mM βOG = 0.35% | 660,000 | [36] |
Human liver | 0.5% Tween 20 | Gel filtration/0.1% Tween 20 | 274,000 | [109] |
Recombinant rat CPT-2 from E. coli | 0.1% Triton X-100/1% βOG | Analytical ultracentrifugation/0.1–1% βOG |
(1) 73,500, 1% βOG (2) 305,000 =80% tetramer, 0.1% βOG |
[32] |
Recombinant rat CPT-2 from E. coli |
(1) 1% βOG (2) 1.5 mM DDM |
Gel filtration, static light scattering/ (1) 1% βOG (2) 1.5 mM DDM |
(1) 88,000, 1% βOG (2) 117,900, 1,5 mM DDM |
Roche, unpublished results |
Recombinant rat CPT-2 from E. coli | None | Gel filtration, light scattering/none | Hexamer–octamer | [31] |
From the data compiled in Table 1 we conclude that the solution molecular weights reported for CPT-1 and CPT-2 are inconsistent, but can be explained by the use of different detergent conditions for characterization of the individual preparations. While predominantly gel filtration has been used as the analytical method for size determination, the experimental conditions, including detergent concentrations and the methods used for enzyme purification, differ substantially.
CPT-2 solubilized with the detergent Triton X-100 below and above its critical micelle-forming concentration showed formation of CPT-2 oligomers. This could indicate the presence of aggregates in preparations with low detergent contents, whereas at high detergent concentrations the dominant species are CPT-2 molecules incorporated into micelles. The average molecular weight of a Triton X-100 micelle is reported to be 90,000 Da, which may lead to a significant overestimation of the molecular weight in solution.
In the authors’ laboratory rat CPT-2 eluted as pure monomer from preparative size-exclusion chromatography at peak concentrations exceeding 10 mg/ml in the presence of 1% (v/v) β-OG or 0.1% (v/v) Triton X-100, while preparations with n-dodecyl-β-d-maltoside or CHAPS still contained oligomers and aggregates. These findings were corroborated by analytical ultracentrifugation and clearly demonstrated that the choice of detergent has a profound effect on the quaternary structure of CPT-2. The CPT-2 construct used for these studies comprised amino acids 27–658, i.e., the mitochondrial leader sequence was removed and replaced by an amino-terminal His-tag comprising 20 residues. Obviously, wild-type CPT-2 is devoid of this amino-terminal (NT) modification when purified from native tissues or expressed with a different tag configuration [31] and we cannot exclude the possibility that this might affect the oligomerization and the membrane attachment of CPT-2. The crystal structures of CPT-2 [31, 32] and our recent structure-based computational model of its membrane association [33] further support that CPT-2 is a monomeric protein. In addition, a query on the PISA server [34] including all deposited CPT-2 crystal structures does not predict the formation of any homo-oligomeric complexes (our unpublished results). Based on the abovementioned data, we propose a monomeric structure for CPT-2.
Besides the oligomerization state, the choice of detergent and the formation of mixed micelles with substrate molecules may also have a pronounced effect on CPT-2 activity [35, 36]. The availability of the substrates palmitoylcarnitine (forward reaction, acyl-import into mitochondria) or palmitoyl-CoA (reverse reaction, acyl-export) is affected by the detergent used for solubilization of CPT-2. Assays with submicellar concentrations of detergents have been described [36], and a well defined lipid environment is likely to be important for maximum activity. The membrane anchor of CPT-2 inserts into the inner leaflet of the inner mitochondrial membrane, approximately one tenth of which consists of the lipid cardiolipin [37], a lipid required for the activity and stability of the transporter component of the CPT system, CACT [38, 39]. Analysis of the crystal structures of CPT-2 and its orientation relative to the membrane based on our previous computational model [33] identified a number of basic residues that could potentially interact with the negatively charged cardiolipin. It would be a great extension of our understanding of the CPT system if structural and enzyme kinetic data of CPT-2 preparations containing cardiolipin became available.
The only systematic study aimed at revealing the quaternary structure of L-CPT-1 was published recently by Faye et al. [40]. This study describes the existence of CPT-1 trimers and their potential to assemble into hexamers. Interestingly, the authors allocate the interaction surface to the NT first 150 residues preceding the catalytic core of CPT-1. This is in agreement with unpublished results from the authors’ laboratory which demonstrated the catalytic cores of both L-CPT-1 and M-CPT-1 to be monomeric in the absence of the first 167 amino acids (Table 1). The NT domain is important for mitochondrial targeting and the malonyl-CoA sensitivity of CPT-1 isoforms [41, 42] and also comprises the two transmembrane helices. The sequence that is necessary and sufficient for the oligomerization of CPT-1 was further mapped to residues 97–147 of the NT domain, as opposed to residues 1–47, which are capable of interacting with the catalytic domain in an intramolecular manner and mediate malonyl-CoA sensitivity [43]. The current model of CPT-1 membrane topology predicts that residues 48–122 cover two transmembrane helices and a connecting loop situated in the mitochondrial intermembrane space. This membrane topology is evolutionarily conserved, because similar positions of the transmembrane helices were proposed on the basis of the sequence of a CPT-1-like protein from the fruit fly Drosophila melanogaster [44].
Faye et al. [40] hypothesized that the cooperativity of CPT-1 with regard to palmitoyl-CoA turnover and malonyl-CoA inhibition may be attributed to the oligomerization of CPT-1.
Overall, investigation of the oligomerization of the isoforms of CPT-1 is currently less complete, but oligomerization could be an important mechanism for regulation of their biological function. Electron microscopy on preparations of intact mitochondrial membranes, including contact sites that are enriched with CPT molecules [45–47], might help to elucidate how the CPT-1 isoforms are assembled.
The crystal structures of CPT-2
The crystal structure of rat CPT-2 was recently elucidated and published independently by two groups [31, 32]. Although solved in different crystallographic space groups and at medium to high resolution (2.6–1.6 Å), the structures are essentially identical as judged from values of 0.23–0.47 Å for root-mean-square-deviations of equivalent Cα-positions [33]. CPT-2 has the same fold as other carnitine acyltransferases with known crystal structures, namely carnitine acetyltransferase (CrAT) [48–52] and carnitine octanoyltransferase [53].
A structurally intriguing aspect of CrAT is its conversion into a “pseudo CPT” by introduction of just two point mutations, D356A and M564D [54]. In addition, the combined mutations Ala106Met/Thr465Val/Thr467Asn/Arg518Asn were found to shift substrate preference from carnitine to choline. These findings emphasize the high degree of structural similarity among carnitine acyltransferases and their similarity to the vital enzyme choline acetyltransferase [55, 56], and need to be exploited for the development of selective inhibitors.
CPT-2, like other carnitine acyltransferases, consists of NT and CT domains, each of which is made up of a six-stranded, central anti-parallel β-sheet surrounded by α-helices (Fig. 3). A common feature of the carnitine acyltransferases is that their NT and CT domains show a significant degree of structural homology within a given enzyme and with chloramphenicol acetyltransferase (CAT) and dihydrolipoyl transacetylase (E2pCD) [48]. CAT and E2pCD are functional trimers with their catalytic sites situated at protomer interfaces, and the active site tunnel of carnitine acyltransferases penetrates these enzymes at the interface of their NT and CT domains. This finding was attributed to a potential gene duplication event and it was hypothesized that a once developed (oligomeric) acetyl-transferase activity could have been transferred to a single polypeptide chain.
Fig. 3.
Positions of post-translational modifications of CPT-1 mapped on the crystal structure of rCPT-2. a Complete view of the structure that shows the modifications to lie in vicinity of the catalytic site tunnel, except for Tyr282 (Tyr162 in rCPT-2). Color scheme: orange loop containing Ser741 and Ser 747 of CPT-1 (627-633 in rCPT-2); cyan: Thr588 and Thr604 of CPT-1 (Thr485 and Thr501 in rCPT-2); yellow Tyr589 of CPT-1 (Tyr486 of rCPT-1), magenta compound 16, green helices, blue sheets, gray loops, NT blue sphere CT red sphere. b Zoom rotated 45° towards point of view and with NT-domain removed for clarity
As yet there is no experimental structure of any of the three known CPT-1 isoforms, but a high degree of sequence homology in the core catalytic domain suggests that their structures are similar to that of CPT-2. Various homology models of CPT-1 have been reported [57–61] and it is certain that the CPT-2 structures have had a significant impact on better understanding of the structural and functional properties of CPT-1.
Mechanism of acylcarnitine and carnitine shuttle in mitochondria
The crystal structures of CPT-2 [31, 32], the location of detergent molecules within these structures, and computational models of the membrane association [33] suggest a membrane association mediated by residues Asn179–Asn208. This insertion is not found in any other type of carnitine acyltransferase with known sequence and contains a pair of anti-parallel helices. The distribution of hydrophobic and polar amino acids renders this insert a likely candidate for the membrane anchor of CPT-2.
Acylcarnitine moieties imported into the mitochondrial matrix by CACT do not equilibrate with the pool of acylcarnitine in the mitochondrial lumen [62]. On the basis of this finding it is intriguing to hypothesize that CPT-2 could be localized on the inner mitochondrial membrane by direct contact with CACT. With the CPT-2-specific insert submerged into the membrane or bound to CACT, the large entrance to the active site via the CoA site of CPT-2 would also be oriented towards the membrane. In this arrangement channeling of acylcarnitine substrates could occur from the transporter CACT into the receiving enzyme CPT-2. A tight juxtaposition of CACT and CPT-2 would generate a microenvironment from which the carnitine that emerges after transesterification could be transported back into the cytosol by CACT.
Another point of interest is the biochemical link of CrAT and CACT. Most hepatic CrAT activity is located in the mitochondrial lumen, where it functions to buffer the concentration of activated acetyl-CoA by equilibration with acetylcarnitine [63]. Excess amounts of acetyl-CoA derived from a high rate of β-oxidation can be transported into the cytosol via CACT. Transfer of the acetyl group from CoA on to carnitine by CrAT is a prerequisite for this process, and it would be interesting to know whether CPT-2 is significantly involved. If it is, then the association of CPT-2 with the membrane and/or its putative interaction with CACT needs to be established in such a way that bidirectional transport of acyl-CoA to and from CPT-2 is possible.
In order to improve our understanding of the stimulation of gluconeogenesis by acetyl-CoA it would be interesting to investigate the import of palmitoylcarnitine and the export of acetylcarnitine. Evidence of simultaneous operation of these processes and information about whether they occur at similar pools of CACT (contact sites vs. bulk membrane) would add significant knowledge.
Post-translational modification of CPT-1 and its correlation with the structure of CPT-2
Hoppel et al. [64–66] identified Ser741 and Ser747, and Thr588 and Thr604, as in-vitro and in-vivo phosphorylation sites, respectively, on L-CPT-1. Phosphorylation of L-CPT-1 on the two carboxy-terminal sites Ser741 and Ser747 was associated with increased catalytic activity and modulation of the response to the physiological inhibitor malonyl-CoA [64].
Residues Ser741 and Ser747 are not conserved between L-CPT-1 and CPT-2. Based on sequence alignments and the structure of rCPT-2 they would presumably be located in an exposed loop, which connects strand β18 with the CT helix α27. As judged from the corresponding region in the rCPT-2 structure, i.e., residues 627–633, this loop would be readily accessible for post-translational modification (Fig. 3).
In contrast, residues Thr588 and Thr604 are homologous with Thr485 and Thr501 in rCPT-2. The likely effect of phosphorylation of these residues can be discussed on the basis of the high sequence homology in the residues surrounding the putative phosphorylation sites. The rCPT-2 crystal structure shows these residues to lie within strands β13 and β14 (with nearly 60% amino acid identity to equivalent residues in L-CPT-1) as part of the extended anti-parallel β-sheet in the core of CPT-2. For steric reasons, this position is difficult to reconcile with their phosphorylation by a conventional protein kinase. One possible explanation was provided in a recent review by Zammit [30], who proposed that phosphorylation of L-CPT-1 may already occur during translation. In contrast with Thr485 of rCPT-2, residue Thr501 has two arginine residues (Arg503 and Arg554) in its vicinity, both of which are fully conserved in the CPT enzyme family and could compensate the two negative charges introduced by phosphorylation. In addition, phosphorylation of Thr501 would certainly change the position of the side chain of Tyr486 (Tyr589 in L-CPT-1), that directly interacts with the carboxyl group of carnitine. It is, therefore, likely that phosphorylation on the equivalent Thr604 of L-CPT-1 would have an effect on affinity for the substrate and catalysis. Obviously, phosphorylation of Thr485 (Thr588 in L-CPT-1) would also change the orientation of the side chain of Tyr486 and, accordingly, the interaction of Tyr589 of L-CPT-1 with the aminocarnitine head-group of substrate or inhibitor molecules would be altered.
If L-CPT-1 is constitutively phosphorylated it would be interesting to know the effect of this modification on the structural integrity of the enzyme, and how L-CPT-1 from native tissue relates to recombinant material in that regard. In a scenario where pools of differentially phosphorylated or otherwise modified (see below) L-CPT-1 exist, it would be essential to examine whether the modifications affect the response of L-CPT-1 to inhibitors.
In addition to phosphorylation, it has been demonstrated that both liver and muscle isoforms of CPT-1 can be nitrated [67–69]. This modification occurs on Tyr282 and on the highly conserved Tyr589 (Tyr486 in rCPT-2; Fig. 3). Nitration of Tyr589 would be likely to affect substrate binding, as discussed above for phosphorylation of neighboring Thr588. The degree of M-CPT-1 nitration in cardiac muscle is increased during sepsis and inflammation; this was hypothesized to underlie the concomitant decrease in M-CPT-1 activity observed for these conditions [69].
L-CPT-1 isolated from rat liver has also been shown to be acetylated on amino-terminal Ala2 [66]. The modifications mentioned above may be part of allosteric or pathological mechanisms for regulating CPT-1 activity and stability.
Substances that modify CPT-activity
Pharmacological modulation of the CPT system has been discussed for treatment of multiple diseases, including T2D, obesity [5], cardiac reperfusion injury (i.e., ischemic crisis) [70] and psoriasis [71]. Treatment of T2D by inhibition of L-CPT-1 has raised special attention, has been examined by several investigators, and was recently reviewed [72–75]. Here, we will focus the discussion on recently published biochemical data and their interpretation in the light of the available crystal structures of CPT-2.
The physiological inhibitor of L-CPT-1 is malonyl-CoA and understanding the details of its interaction with L-CPT-1 may reveal how the regulatory NT-domain interferes with the activity of the catalytic domain.
Interestingly, the mutation of the active site residue Met593 to Ser, Ala and Glu, almost abrogate inhibition by malonyl-CoA [76]. The importance of this region is confirmed by the finding that the intrinsically less malonyl-CoA-sensitive isoform L-CPT-1 displays a marked increase in inhibition by malonyl-CoA on introduction of a single point mutation, Glu590Ala [77]. Both Glu590 and Met593 are part of strand β13, which lines the entrance to the active site via the CoA site. Mutation of Glu590 also destroys an internal salt-bridge to the highly conserved His640 of helix α23 (His539 in rat CPT2), which might affect the structural integrity of this region of the enzymatic core.
These results are corroborated by mutation analysis of residues located in the amino-terminal regulatory domain. The point mutations Glu3Ala, Ser24Ala, and Gln30Ala diminish the interaction with the catalytic core of L-CPT-1 [43]. In addition, an increase in membrane fluidity of the mitochondrial outer membrane—which was demonstrated to occur because of a change in lipid composition in diabetic rats—disrupts the interaction of the NT and CT domains [43]. The available data imply a high degree of conformational flexibility and complex mutual interactions for the NT and CT domains, which would certainly complicate drug-discovery programs targeting the allosteric malonyl-CoA site. This view is supported by attempts in the authors’ laboratory to co-express the soluble parts of the NT and CT domains of CPT-1, which did not yield a stable complex (unpublished results). Nevertheless, mutagenesis studies in combination with in-silico modeling have been instrumental in identifying the contact sites of the regulatory amino-terminal and carboxy-terminal catalytic domains of CPT-1, both with regard to each other and to the binding site of malonyl-CoA [43, 61, 76, 78, 79].
So far, the α-ketoacid compound Ro25-0187 is the only inhibitor whose binding site has been mapped to the regulatory malonyl-CoA site by competition experiments [80, 81]. Despite the presence of an allosteric mechanism of inhibition for CPT-1 by malonyl-CoA [82], drug-discovery programs have devoted their effort to the catalytic site. This may turn out to be the more feasible route, because crystal structures of CPT-2 have become available to guide compound development whereas no experimental structural information is yet available about the regulatory domain of CPT-1 enzymes. Moreover, it is assumed that the deep tripartite active site tunnel with pronounced hydrophobic and polar features has higher drugability than the allosteric site, which was modeled to lie at a rather shallow interface of the regulatory and catalytic domains [61].
With the isolation of l-aminocarnitine derivatives, for example the emericedins and similar substrate analogs, potent catalytic site inhibitors of CPT enzymes became available [11–17]. These studies demonstrated the anti-diabetic effects of some acylaminocarnitine compounds and their metabolically more stable and non-hydrolyzable urea derivatives [15, 16]. The strongest hypoglycemic and anti-ketogenic effects were observed with palmitoyl-substituted compounds, which were most similar to the preferred in vivo CPT substrate, palmitoylcarnitine. These compounds do not show isoform specificity and l-aminocarnitine is able to inhibit the ubiquitous CPT-2 in intact mitochondria [83]. However, substrate analogs are valuable research tools for early target validation and proof of concept investigations.
The compound class glycidic acid derivatives [84, 85] has been characterized as covalent and irreversible CPT inhibitors, as shown for the best studied members of this class, tetradecyl glycidic acid (TDGA) and etomoxir. Cardiotoxic side effects of glycidic acid derivatives have been attributed to high reactivity of their epoxide moiety [74]. It has been stated that etomoxir (and 2-oxiranecarboxylic acids in general) have to be converted to CoA esters to become potent inhibitors of CPT-1 [86]. In this sense etomoxir is a precursor of an inhibitor.
Phenylalkyl oxirane carboxylates (POCA) exert their full efficacy as CPT-1 inhibitors when converted to POCA-CoA [87–89]. Besides inhibiting CPT-1, POCA stimulates peroxisome proliferation similar to fibrates and thereby increases lipid catabolism via peroxisomal oxidation.
The best investigated small-molecule inhibitors of CPT enzymes are ST1326 and etomoxir (Fig. 4), and a rich source of literature is available on the in-vitro and in-vivo pharmacology of these compounds. ST1326 is a non-cleavable analog of palmitoylcarnitine, the physiological substrate of CPT-2. The crystal structure of CPT-2 with ST1326 [32] located the binding interface to the acyl and carnitine tunnels of the active site of CPT-2 and furnished a detailed view of the binding mode. Hydrophobic interaction of the aliphatic 14 carbon atom chain contributes substantially to the binding affinity of ST1326; this is further supported by detection of electron density in the apo-structure that can be interpreted as an aliphatic C16 chain at this location [31]. The more hydrophilic aminocarnitine head group contributes three direct hydrogen bonds to the amino acid side chains of His372, Ser590, and Tyr486 and one bond to a water molecule, whereas no covalent interaction could be observed. In contrast, no crystal structure of the irreversible covalent inhibitor etomoxir has yet been determined. However, the current data support the view that the activated Ser590 is positioned well to provide the covalent bond. Similarities in the structure (Fig. 4) with regard to the aliphatic carbon chain suggest a similar position of this part of the molecule compared with ST1326.
Fig. 4.
Ligands that modify the activity of CPT enzymes. The active form of etomoxir is assumed to be its CoA derivative (see text for details)
Assessment of the CPT system as drug target
The biological mechanism of the CPT system is an attractive target for treatment of T2D, on the basis of its key role in maintaining fatty acid metabolism and its concomitant effect on glucose homeostasis. The proof of concept for triglyceride-lowering and hypoglycemic effects of the CPT inhibitors etomoxir and ST1326 have been demonstrated in humans and add this enzyme to the targets for treatment of T2D.
In addition, tools for a successful drug-discovery program are now available, such as high quality protein preparations, robust assays, and, recently, the structure of CPT-2. A number of potent inhibitors of CPT enzymes and crystal structures of complexes with these compounds show the good chemical tractability of the modulation of CPT activity by small molecules. The availability of compounds for the modulation of enzyme activity together with biochemical and in-vivo data provides an excellent basis for discussing of the consequences of the drug-mediated modulation of the CPT system.
The key open point that must be addressed is the optimal isoform specificity of inhibitors of the CPT system. For example, examination of rat cardiac myocytes is difficult in that they express both M-CPT-1 and L-CPT-1 isoforms, until an isoform switch occurs during the early postnatal phase, after which M-CPT-1 remains the predominant isoform in the rat heart [90–92]. The cardiac expression profile of CPT-1 isoforms in humans needs further investigation in order to improve the characterization of cardiotoxic site effects.
The detrimental effects of a non-specific inhibitor are exemplified by accumulation of ectopic lipids in skeletal muscle of rats after prolonged oral administration of etomoxir, which inhibits both L-CPT-1 and M-CPT-1 irreversibly [111]. Merrill et al. [93] described high-dose (1 mM) in-vitro studies with etomoxir that demonstrated induction of acute oxidative stress in HepG2 cells, as exemplified by upregulation of glutathione reductase and thioredoxin reductase mRNA after 6 h of acute treatment. Reduced levels of glutathione as a consequence of oxidative stress induced by etomoxir and the reversible CPT-1 inhibitor CPI975 were also observed for tissue slices of rat and human livers [94]. Besides these changes in cellular redox-status, abnormalities in gene expression profiles and mitochondrial morphology were reported. Although even high doses of CPI975 do not induce cardiac hypertrophy as observed with etomoxir, a significant body of evidence emphasizes the hepatotoxic effect of CPI975 [74]. The exact mechanism that couples L-CPT-1 inhibition with the occurrence of mitochondrial lesions remains to be elucidated. The side effects of etomoxir and related CPT inhibitors derived from glycidic acid may not be mechanism-based but mediated by the reactivity of the epoxide moiety of these compounds. Complete inhibition, as in the case of irreversible inhibitors such as etomoxir, should clearly be avoided in long-term pharmacological treatment because this could provoke the accumulation of lipids in the cytosol of cells other than adipocytes and deplete the mitochondrial matrix of precursors for membrane lipids.
When substrate flux through the CPT system to β-oxidation is reduced by inhibition of mitochondrial CPT-1, metabolism is shifted toward utilization of glucose, as described by Randle [95, 96]. Besides being an advantage for ameliorating the symptoms of T2D, this metabolic switch can be exploited for treatment of stable angina pectoris and myocardial crisis. Carbohydrate oxidation needs less oxygen than oxidation of fatty acids for equivalent energy supplies. Two excellent and recent reviews [97, 98] summarize use of the CPT inhibitor perhexiline as antianginal drug. The occurrence of hepatotoxic and neurotoxic side effects under prolonged perhexiline treatment has been attributed to its lack of CPT isoform specificity.
However, two compounds have been described that have some isoform selectivity, although these inhibitors achieve their specificity by different means. In cardiac myocytes, and to a lesser extend in hepatocytes, the compound oxfenicine (S-2-(4-hydroxyphenyl)glycine) is converted to hydroxyphenylglyoxylate, which is an inhibitor of CPT-1 with lower IC50 for M-CPT-1 [99]. The conversion is catalyzed by a mitochondrial aminotransferase that has higher activity in heart muscle than in the liver. Together these two effects result in marked tissue specificity. Another example is a dinitrophenol derivative of etomoxir (DNP–etomoxir) which is a potent and specific inhibitor of L-CPT-1 [68, 91].
Several reports have documented the effect of CPT-1 inhibitors in the central nervous system, especially the hypothalamus, and have pointed out the exquisite role of the hypothalamus–pituitary axis and its interaction with the periphery via peptide hormones in whole-body energy homeostasis [10, 100–102].
Knock-down and pharmacological inhibition of CPT-1 activity in rat hypothalamus has been shown to result in reduced food intake and reduction of endogenous glucose production [10]. Interestingly, in this study genetic abrogation of L-CPT-1 activity was sufficient to enable observation of these effects. Therefore, the brain-specific isoform B-CPT-1, although expressed in hypothalamic neurons [6], does not seem to be involved in this process, even though B-CPT-1 KO mice show a similar phenotype of hypophagia [7].
Therefore, the currently available data suggest that a drug molecule targeted at the CPT system ought to be isoform-specific and be distributed non-systemically to avoid mechanism-related side effects. This leaves the individual CPT-1 isoforms that are differentially expressed in the relevant tissues as drug targets, because CACT and CPT-2 are ubiquitously expressed enzymes. An alternative option might be a compound lacking isoform selectivity but with pronounced organ-specificity, e.g., because of specific transport processes. Drug–drug interactions must be considered, because of the frequent clinical coherence of T2D and obesity or dyslipidemia as symptoms of the metabolic syndrome. As a very special case, the concomitant administration of CPT inhibitors and statins was investigated with regard to aggravation of the very rare sub-symptomatic CPT-2 deficiency into an overt form [103–105].
With all the research tools of modern drug discovery in hand, the discovery and development of chemical entities that modulate the activity of the CPT, safely and with high efficacy, for treatment of patients, is within reach.
Acknowledgments
We thank Simona Ceccarelli-Grenz and Odile Chomienne for critical reading of the manuscript.
References
- 1.ADA Report of the expert committee on the diagnosis and classification of diabetes mellitus. Diabetes Care. 2003;26:5–20. doi: 10.2337/diacare.26.2007.S5. [DOI] [PubMed] [Google Scholar]
- 2.Zimmet P, Alberti KG, Shaw J. Global and societal implications of the diabetes epidemic. Nature. 2001;414:782–787. doi: 10.1038/414782a. [DOI] [PubMed] [Google Scholar]
- 3.Wagman AS, Nuss JM. Current therapies and emerging targets for the treatment of diabetes. Curr Pharm Des. 2001;7:417–450. doi: 10.2174/1381612013397915. [DOI] [PubMed] [Google Scholar]
- 4.Nathan DM. Rosiglitazone and cardiotoxicity—weighing the evidence. N Engl J Med. 2007;357:64–66. doi: 10.1056/NEJMe078117. [DOI] [PubMed] [Google Scholar]
- 5.Kuhajda FP, Ronnett GV. Modulation of carnitine palmitoyltransferase-1 for the treatment of obesity. Curr Opin Inves Drugs. 2007;8:312–317. [PubMed] [Google Scholar]
- 6.Price N, van der Leij F, Jackson V, Corstorphine C, Thomson R, Sorensen A, Zammit V. A novel brain-expressed protein related to carnitine palmitoyltransferase I. Genomics. 2002;80:433–442. doi: 10.1006/geno.2002.6845. [DOI] [PubMed] [Google Scholar]
- 7.Wolfgang MJ, Kurama T, Dai Y, Suwa A, Asaumi M, Matsumoto S, Cha SH, Shimokawa T, Lane MD. The brain-specific carnitine palmitoyltransferase-1c regulates energy homeostasis. Proc Natl Acad Sci USA. 2006;103:7282–7287. doi: 10.1073/pnas.0602205103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Sierra AY, Gratacós E, Carrasco P, Clotet J, Ureña J, Serra D, Asins G, Hegardt FG, Casals N. CPT1c is localized in endoplasmic reticulum of neurons and has carnitine palmitoyltransferase activity. J Biol Chem. 2008;283:6878–6885. doi: 10.1074/jbc.M707965200. [DOI] [PubMed] [Google Scholar]
- 9.Roomets E, Kivelä T, Tyni T. Carnitine palmitoyltransferase I and acyl-CoA dehydrogenase 9 in retina: insights of retinopathy in mitochondrial trifunctional protein defects. Invest Ophthalmol Vis Sci. 2008;49:1660–1664. doi: 10.1167/iovs.07-1094. [DOI] [PubMed] [Google Scholar]
- 10.Obici S, Feng Z, Arduini A, Conti R, Rossetti L. Inhibition of hyopothalmic carnitine palmitoyltransferase-1 decreases food intake and glucose production. Nat Med. 2003;6:756–761. doi: 10.1038/nm873. [DOI] [PubMed] [Google Scholar]
- 11.Kanamaru T, Shinagawa S, Asai M, Okazaki H, Sugiyama Y, Fujita T, Iwatsuka H, Yoneda M. Emeriamine, an antidiabetic beta-aminobetaine derived from a novel fungal metabolite. Life Sci. 1985;37:217–223. doi: 10.1016/0024-3205(85)90647-2. [DOI] [PubMed] [Google Scholar]
- 12.Shinagawa S, Kanamaru T, Harada S, Asai M, Okazaki H. Chemistry and inhibitory activity of long chain fatty acid oxidation of emeriamine and its analogues. J Med Chem. 1987;30:1458–1463. doi: 10.1021/jm00391a030. [DOI] [PubMed] [Google Scholar]
- 13.Jenkins DL, Griffith OW. dl-aminocarnitine and acetyl-dl-aminocarnitine. Potent inhibitors of carnitine acyltransferases and hepatic triglyceride catabolism. J Biol Chem. 1985;260:14748–14755. [PubMed] [Google Scholar]
- 14.Jenkins DL, Griffith OW. Antiketogenic and hypoglycemic effects of aminocarnitine and acylaminocarnitines. Proc Natl Acad Sci USA. 1986;83:290–294. doi: 10.1073/pnas.83.2.290. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Giannessi F, Chiodi P, Marzi M, Minetti P, Pessotto P, De Angelis F, Tassoni E, Conti R, Giorgi F, Mabilia M, Dell’Uomo N, Muck S, Tinti MO, Carminati P, Arduini A. Reversible carnitine palmitoyltransferase inhibitors with broad chemical diversity as potential antidiabetic agents. J Med Chem. 2001;44:2383–2386. doi: 10.1021/jm010889+. [DOI] [PubMed] [Google Scholar]
- 16.Giannessi F, Pessotto P, Tassoni E, Chiodi P, Conti R, De Angelis F, Dell’Uomo N, Catini R, Deias R, Tinti MO, Carminati P, Arduini A. Discovery of a long-chain carbamoyl aminocarnitine derivative, a reversible carnitine palmitoyltransferase inhibitor with antiketotic and antidiabetic activity. J Med Chem. 2003;46:303–309. doi: 10.1021/jm020979u. [DOI] [PubMed] [Google Scholar]
- 17.Chegary M, Te Brinke H, Doolaard M, Ijlst L, Wijburg FA, Wanders RJ, Houten SM. Characterization of l-aminocarnitine, an inhibitor of fatty acid oxidation. Mol Genet Metab. 2008;93:403–410. doi: 10.1016/j.ymgme.2007.11.001. [DOI] [PubMed] [Google Scholar]
- 18.Fritz IB, Yue KT. Long-chain carnitine acyltransferase and the role of acylcarnitine derivatives in the catalytic increase of fatty acid oxidation induced by carnitine. J Lipid Res. 1963;4:279–288. [PubMed] [Google Scholar]
- 19.Esser V, Kuwajima M, Britton CH, Krishnan K, Foster DW, McGarry JD. Inhibitors of mitochondrial carnitine palmitoyltransferase I limit the action of proteases on the enzyme. Isolation and partial amino acid analysis of a truncated form of the rat liver isozyme. J Biol Chem. 1993;268:5810–5816. [PubMed] [Google Scholar]
- 20.Esser V, Britton CH, Weis BC, Foster DW, McGarry JD. Cloning, sequencing, and expression of a cDNA encoding rat liver carnitine palmitoyltransferase I. Direct evidence that a single polypeptide is involved in inhibitor interaction and catalytic function. J Biol Chem. 1993;268:5817–5822. [PubMed] [Google Scholar]
- 21.Yates DW, Garland PB. Carnitine palmitoyltransferase activities (EC 2.3.1.-) of rat liver mitochondria. Biochem J. 1970;119:547–552. doi: 10.1042/bj1190547. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Kopec B, Fritz IB. Properties of a purified carnitine palmitoyltransferase, and evidence for the existence of other carnitine acyltransferases. Can J Biochem. 1971;49:941–948. doi: 10.1139/o71-136. [DOI] [PubMed] [Google Scholar]
- 23.West DW, Chase JFA, Tubbs PK. The separation and properties of two forms of carnitine palmitoyltransferase from ox liver mitochondria. Biochem Biophys Res Comm. 1971;42:912–918. doi: 10.1016/0006-291X(71)90517-1. [DOI] [PubMed] [Google Scholar]
- 24.Solberg HE. Different carnitine acyltransferases in calf liver. Biochem Biophys Acta. 1972;280:422–433. doi: 10.1016/0005-2760(72)90248-2. [DOI] [PubMed] [Google Scholar]
- 25.Kopec B, Fritz IB. Comparison of properties of carnitine palmitoyltransferase I with those of carnitine palmitoyltransferase II, and preparation of antibodies to carnitine palmitoyltransferases. J Biol Chem. 1973;248:4069–4074. [PubMed] [Google Scholar]
- 26.Brown NF, Sen A, Soltis DA, Jones B, Foster DW, McGarry JD. Expression of precursor and mature carnitine palmitoyltransferase II in Escherichia coli and in vitro: differential behaviour of rat and human isoforms. Biochem J. 1993;294:79–86. doi: 10.1042/bj2940079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.McGarry JD, Brown NF. Reconstitution of purified, active and malonyl-CoA-sensitive rat liver carnitine palmitoyltransferase I: relationship between membrane environment and malonyl-CoA sensitivity. Biochem J. 2000;349:179–187. doi: 10.1042/0264-6021:3490179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Brown NF, Esser V, Foster DW, McGarry JD. Expression of a cDNA for rat liver carnitine palmitoyltransferase I in yeast establishes that catalytic activity and malonyl-CoA sensitivity reside in a single polypeptide. J Biol Chem. 1994;269:26438–26442. [PubMed] [Google Scholar]
- 29.Prip-Buus C, Cohen I, Kohl C, Esser V, McGarry JD, Girard J. Topological and functional analysis of the rat liver carnitine palmitoyltransferase 1 expressed in Saccharomyces cerevisiae . FEBS Lett. 1998;429:173–178. doi: 10.1016/S0014-5793(98)00584-5. [DOI] [PubMed] [Google Scholar]
- 30.Zammit VA. Carnitine palmitoyltransferase 1: central to cell function. IUBMB Life. 2008;60:347–354. doi: 10.1002/iub.78. [DOI] [PubMed] [Google Scholar]
- 31.Hsiao YS, Jogl G, Esser V, Tong L. Crystal structure of rat carnitine palmitoyltransferase II (CPT-II) Biochem Biophys Res Commun. 2006;346:974–980. doi: 10.1016/j.bbrc.2006.06.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Rufer AC, Thoma R, Benz J, Stihle M, Gsell B, De Roo E, Banner DW, Mueller F, Chomienne O, Hennig M. The crystal structure of carnitine palmitoyltransferase 2 and implications for diabetes treatment. Structure. 2006;14:713–723. doi: 10.1016/j.str.2006.01.008. [DOI] [PubMed] [Google Scholar]
- 33.Rufer AC, Lomize A, Benz J, Chomienne O, Thoma R, Hennig M. Carnitine palmitoyltransferase 2: analysis of membrane association and complex structure with a substrate analog. FEBS Lett. 2007;581:3247–3252. doi: 10.1016/j.febslet.2007.05.080. [DOI] [PubMed] [Google Scholar]
- 34.Krissinel E, Henrick K. Inference of macromolecular assemblies from crystalline state. J Mol Biol. 2007;372:774–797. doi: 10.1016/j.jmb.2007.05.022. [DOI] [PubMed] [Google Scholar]
- 35.Clarke PR, Bieber LL. Isolation and purification of mitochondrial carnitine octanoyltransferase activities from beef heart. J Biol Chem. 1981;256:9861–9868. [PubMed] [Google Scholar]
- 36.Fiol CJ, Bieber LL. Effects of octylglucoside and triton X-100 on the kinetics and specificity of carnitine palmitoyltransferase. Lipids. 1988;23:120–125. doi: 10.1007/BF02535291. [DOI] [PubMed] [Google Scholar]
- 37.Nichols-Smith S, Kuhl T. Electrostatic interactions between model mitochondrial membranes. Colloids Surf B Biointerfaces. 2005;41:121–127. doi: 10.1016/j.colsurfb.2004.11.003. [DOI] [PubMed] [Google Scholar]
- 38.Rubio-Gozalbo ME, Bakker JA, Waterham HR, Wanders RJ. Carnitine–acylcarnitine translocase deficiency, clinical, biochemical and genetic aspects. Mol Aspects Med. 2004;25:521–532. doi: 10.1016/j.mam.2004.06.007. [DOI] [PubMed] [Google Scholar]
- 39.Palsdottir H, Hunte C. Lipids in membrane protein structures. Biochim Biophys Acta. 2004;1666:2–18. doi: 10.1016/j.bbamem.2004.06.012. [DOI] [PubMed] [Google Scholar]
- 40.Faye A, Esnous C, Price NT, Onfray MA, Girard J, Prip-Buus C. Rat liver carnitine palmitoyltransferase 1 forms an oligomeric complex within the outer mitochondrial membrane. J Biol Chem. 2007;282:26908–26916. doi: 10.1074/jbc.M705418200. [DOI] [PubMed] [Google Scholar]
- 41.Cohen I, Kohl C, McGarry JD, Girard J, Prip-Buus C. The N-terminal domain of rat liver carnitine palmitoyltransferase 1 mediates import into the outer mitochondrial membrane and is essential for activity and malonyl-CoA sensitivity. J Biol Chem. 1998;273:29896–29904. doi: 10.1074/jbc.273.45.29896. [DOI] [PubMed] [Google Scholar]
- 42.Cohen I, Guillerault F, Girard J, Prip-Buus C. The N-terminal domain of rat liver carnitine palmitoyltransferase 1 contains an internal mitochondrial import signal and residues essential for folding of its C-terminal catalytic domain. J Biol Chem. 2001;276:5403–5411. doi: 10.1074/jbc.M009555200. [DOI] [PubMed] [Google Scholar]
- 43.Faye A, Borthwick K, Esnous C, Price NT, Gobin S, Jackson VN, Zammit VA, Girard J, Prip-Buus C. Demonstration of N- and C-terminal domain intramolecular interactions in rat liver carnitine palmitoyltransferase 1 that determine its degree of malonyl-CoA sensitivity. Biochem J. 2005;387:67–76. doi: 10.1042/BJ20041533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Jackson VN, Cameron JM, Zammit VA, Price NT. Sequencing and functional expression of the malonyl-CoA-sensitive carnitine palmitoyltransferase from Drosophila melanogaster . Biochem J. 1999;341:483–489. doi: 10.1042/0264-6021:3410483. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Fraser F, Zammit VA. Enrichment of carnitine palmitoyltransferases I and II in the contact sites of rat liver mitochondria. Biochem J. 1998;329:225–229. doi: 10.1042/bj3290225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Fraser F, Zammit VA. Submitochondrial and subcellular distributions of the carnitine–acylcarnitine carrier. FEBS Lett. 1999;445:41–44. doi: 10.1016/S0014-5793(99)00095-2. [DOI] [PubMed] [Google Scholar]
- 47.Fraser F, Padovese R, Zammit VA. Distinct kinetics of carnitine palmitoyltransferase i in contact sites and outer membranes of rat liver mitochondria. J Biol Chem. 2001;276:20182–20185. doi: 10.1074/jbc.M101078200. [DOI] [PubMed] [Google Scholar]
- 48.Jogl G, Tong L. Crystal structure of carnitine acetyltransferase and implications for the catalytic mechanism and fatty acid transport. Cell. 2003;112:113–122. doi: 10.1016/S0092-8674(02)01228-X. [DOI] [PubMed] [Google Scholar]
- 49.Wu D, Govindasamy L, Lian W, Gu Y, Kukar T, Agbandje-McKenna M, McKenna R. Structure of human carnitine acetyltransferase. Molecular basis for fatty acyl transfer. J Biol Chem. 2003;278:13159–13165. doi: 10.1074/jbc.M212356200. [DOI] [PubMed] [Google Scholar]
- 50.Hsiao YS, Jogl G, Tong L. Structural and biochemical studies of the substrate selectivity of carnitine acetyltransferase. J Biol Chem. 2004;279:31584–31589. doi: 10.1074/jbc.M403484200. [DOI] [PubMed] [Google Scholar]
- 51.Govindasamy L, Kukar T, Lian W, Pedersen B, Gu Y, Agbandje-McKenna M, Jin S, McKenna R, Wu D. Structural and mutational characterization of l-carnitine binding to human carnitine acetyltransferase. J Struct Biol. 2004;146:416–424. doi: 10.1016/j.jsb.2004.01.011. [DOI] [PubMed] [Google Scholar]
- 52.Hsiao YS, Jogl G, Tong L. Crystal structures of murine carnitine acetyltransferase in ternary complexes with its substrates. J Biol Chem. 2006;281:28480–28487. doi: 10.1074/jbc.M602622200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Hsiao YS, Tong L. Crystal structure of mouse carnitine octanoyltransferase and molecular determinants of substrate selectivity. J Biol Chem. 2004;280:738–744. doi: 10.1074/jbc.M409894200. [DOI] [PubMed] [Google Scholar]
- 54.Cordente AG, López-Viñas E, Vázquez MI, Gómez-Puertas P, Asins G, Serra D, Hegardt FG. Mutagenesis of specific amino acids converts carnitine acetyltransferase into carnitine palmitoyltransferase. Biochemistry. 2006;45:6133–6141. doi: 10.1021/bi0602664. [DOI] [PubMed] [Google Scholar]
- 55.Cai Y, Cronin CN, Engel AG, Ohno K, Hersh LB, Rodgers DW. Choline acetyltransferase structure reveals distribution of mutations that cause motor disorders. EMBO J. 2004;23:2047–2058. doi: 10.1038/sj.emboj.7600221. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Govindasamy L, Pedersen B, Lian W, Kukar T, Gu Y, Jin S, Agbandje-McKenna M, Wu D, McKenna R. Structural insights and functional implications of choline acetyltransferase. J Struct Biol. 2004;148:226–235. doi: 10.1016/j.jsb.2004.06.005. [DOI] [PubMed] [Google Scholar]
- 57.Morillas M, Gómez-Puertas P, Roca R, Serra D, Asins G, Valencia A, Hegardt FG. Structural model of the catalytic core of carnitine palmitoyltransferase I and carnitine octanoyltransferase (COT): mutation of CPT I histidine 473 and alanine 381 and COT alanine 238 impairs the catalytic activity. J Biol Chem. 2001;276:45001–45008. doi: 10.1074/jbc.M106920200. [DOI] [PubMed] [Google Scholar]
- 58.Morillas M, Gómez-Puertas P, Rubí B, Clotet J, Ariño J, Valencia A, Hegardt FG, Serra D, Asins G. Structural model of a malonyl-CoA-binding site of carnitine octanoyltransferase and carnitine palmitoyltransferase I: mutational analysis of a malonyl-CoA affinity domain. J Biol Chem. 2002;277:11473–11480. doi: 10.1074/jbc.M111628200. [DOI] [PubMed] [Google Scholar]
- 59.Gobin S, Thuillier L, Jogl G, Faye A, Tong L, Chi M, Bonnefont JP, Girard J, Prip-Buus C. Functional and structural basis of carnitine palmitoyltransferase 1A deficiency. J Biol Chem. 2003;278:50428–50434. doi: 10.1074/jbc.M310130200. [DOI] [PubMed] [Google Scholar]
- 60.Woldegiorgis G, Dai J, Arvidson D. Structure–function studies with the mitochondrial carnitine palmitoyltransferase I and II. Chemical Monthly. 2005;136:1325–1340. doi: 10.1007/s00706-005-0334-7. [DOI] [Google Scholar]
- 61.López-Viñas E, Bentebibel A, Gurunathan C, Morillas M, de Arriaga D, Serra D, Asins G, Hegardt FG, Gómez-Puertas P. Definition by functional and structural analysis of two malonyl-CoA sites in carnitine palmitoyltransferase 1A. J Biol Chem. 2007;282:18212–18224. doi: 10.1074/jbc.M700885200. [DOI] [PubMed] [Google Scholar]
- 62.Murthy MS, Pande SV. Microcompartmentation of transported carnitine, acetylcarnitine and ADP occurs in the mitochondrial matrix. Implications for transport measurements and metabolism. Biochem J. 1985;230:657–663. doi: 10.1042/bj2300657. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Ramsay RR, Naismith JH. A snapshot of carnitine acetyltransferase. Trends Biochem Sci. 2003;28:343–346. doi: 10.1016/S0968-0004(03)00137-3. [DOI] [PubMed] [Google Scholar]
- 64.Kerner J, Distler AM, Minkler P, Parland W, Peterman SM, Hoppel CL. Phosphorylation of rat liver mitochondrial carnitine palmitoyltransferase-I: effect on the kinetic properties of the enzyme. J Biol Chem. 2004;279:41104–41113. doi: 10.1074/jbc.M406570200. [DOI] [PubMed] [Google Scholar]
- 65.Kerner J, Hoppel CL. Carnitine palmitoyltransferase-I and regulation of mitochondrial fatty acid oxidation. Chemical Monthly. 2005;136:1311–1323. doi: 10.1007/s00706-005-0329-4. [DOI] [Google Scholar]
- 66.Distler AM, Kerner J, Hoppel CL. Post-translational modifications of rat liver mitochondrial outer membrane proteins identified by mass spectrometry. Biochim Biophys Acta. 2007;1774:628–636. doi: 10.1016/j.bbapap.2007.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Fukumoto K, Pierro A, Spitz L, Eaton S. Differential effects of neonatal endotoxemia on heart and kidney carnitine palmitoyl transferase I. J Pediatr Surg. 2002;37:723–726. doi: 10.1053/jpsu.2002.32263. [DOI] [PubMed] [Google Scholar]
- 68.Fukumoto K, Pierro A, Zammit VA, Spitz L, Eaton S. Tyrosine nitration of carnitine palmitoyl transferase I during endotoxaemia in suckling rats. Biochim Biophys Acta. 2004;1683:1–6. doi: 10.1016/j.bbalip.2004.03.006. [DOI] [PubMed] [Google Scholar]
- 69.Eaton S, Fukumoto K, Stefanutti G, Spitz L, Zammit VA, Pierro A. Myocardial carnitine palmitoyltransferase I as a target for oxidative modification in inflammation and sepsis. Biochem Soc Trans. 2003;31:1133–1136. doi: 10.1042/BST0311133. [DOI] [PubMed] [Google Scholar]
- 70.Lopaschuk GD, Wall SR, Olley PM, Davies NJ. Etomoxir, a carnitine palmitoyltransferase I inhibitor, protects hearts from fatty acid-induced ischemic injury independent of changes in long chain acylcarnitine. Circ Res. 1988;63:1036–1043. doi: 10.1161/01.res.63.6.1036. [DOI] [PubMed] [Google Scholar]
- 71.Caspary F, Elliott G, Navé BT, Verzaal P, Rohrbach M, Das PK, Nagelkerken L, Nieland JD. A new therapeutic approach to treat psoriasis by inhibition of fatty acid oxidation by etomoxir. Br J Dermatol. 2005;153:937–944. doi: 10.1111/j.1365-2133.2005.06811.x. [DOI] [PubMed] [Google Scholar]
- 72.Anderson RC, Balestra M, Bell PA, Deems RO, Fillers WS, Foley JE, Fraser JD, Mann WR, Rudin M, Villhauer EB. Antidiabetic agents: a new class of reversible carnitine palmitoyltransferase I inhibitors. J Med Chem. 1995;38:3448–3450. doi: 10.1021/jm00018a003. [DOI] [PubMed] [Google Scholar]
- 73.Anderson RC. Carnitine palmitoyltransferase: a viable target for the treatment of NIDDM? Curr Pharm Des. 1998;4:1–16. [PubMed] [Google Scholar]
- 74.Kurukulasuriya R, Link JT, Madar DJ, Pei Z, Richards SJ, Rohde JJ, Souers AJ, Szczepankiewicz BG. Potential drug targets and progress towards pharmacologic inhibition of hepatic glucose production. Curr Med Chem. 2003;10:123–153. doi: 10.2174/0929867033368556. [DOI] [PubMed] [Google Scholar]
- 75.Grynberg A. Effectors of fatty acid oxidation reduction: promising new anti-ischaemic agents. Curr Pharm Des. 2005;11:489–509. doi: 10.2174/1381612053382061. [DOI] [PubMed] [Google Scholar]
- 76.Morillas M, Gómez-Puertas P, Bentebibel A, Sellés E, Casals N, Valencia A, Hegardt FG, Asins G, Serra D. Identification of conserved amino acid residues in rat liver carnitine palmitoyltransferase I critical for malonyl-CoA inhibition. Mutation of methionine 593 abolishes malonyl-CoA inhibition. J Biol Chem. 2003;278:9058–9063. doi: 10.1074/jbc.M209999200. [DOI] [PubMed] [Google Scholar]
- 77.Napal L, Dai J, Treber M, Haro D, Marrero PF, Woldegiorgis G. A single amino acid change (substitution of the conserved Glu-590 with alanine) in the C-terminal domain of rat liver carnitine palmitoyltransferase I increases its malonyl-CoA sensitivity close to that observed with the muscle isoform of the enzyme. J Biol Chem. 2003;278:34084–34089. doi: 10.1074/jbc.M305826200. [DOI] [PubMed] [Google Scholar]
- 78.Swanson ST, Foster DW, McGarry JD, Brown NF. Roles of the N- and C-terminal domains of carnitine palmitoyltransferase I isoforms in malonyl-CoA sensitivity of the enzymes: insights from expression of chimaeric proteins and mutation of conserved histidine residues. Biochem J. 1998;335:513–519. doi: 10.1042/bj3350513. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Jackson VN, Cameron JM, Fraser F, Zammit VA, Price NT. Use of six chimeric proteins to investigate the role of intramolecular interactions in determining the kinetics of carnitine palmitoyltransferase I isoforms. J Biol Chem. 2000;275:19560–19566. doi: 10.1074/jbc.M002177200. [DOI] [PubMed] [Google Scholar]
- 80.Cook GA, Mynatt RL, Kashfi K. Yonetani–Theorell analysis of hepatic carnitine palmitoyltransferase-I inhibition indicates two distinct inhibitory binding sites. J Biol Chem. 1994;269:8803–8807. [PubMed] [Google Scholar]
- 81.Yang N, Kays JS, Skillman TR, Burris L, Seng TW, Hammond C. C75 [4-methylene-2-octyl-5-oxo-tetrahydro-furan-3-carboxylic acid] activates carnitine palmitoyltransferase-1 in isolated mitochondria and intact cells without displacement of bound malonyl CoA. J Pharmacol Exp Ther. 2005;312:127–133. doi: 10.1124/jpet.104.074104. [DOI] [PubMed] [Google Scholar]
- 82.McGarry JD, Leatherman GF, Foster DW. Carnitine palmitoyltransferase I. The site of inhibition of hepatic fatty acid oxidation by malonyl-CoA. J Biol Chem. 1978;253:4128–4136. [PubMed] [Google Scholar]
- 83.Traufeller K, Gellerich FN, Zierz S. Different sensitivities of CPT I and CPT II for inhibition by l-aminocarnitine in human skeletal muscle. Biochim Biophys Acta. 2004;1608:149–154. doi: 10.1016/j.bbabio.2003.11.005. [DOI] [PubMed] [Google Scholar]
- 84.Tutwiler GF, Kirsch T, Mohrbacher RJ, Ho W. Pharmacologic profile of methyl 2-tetradecylglycidate (McN-3716)—an orally effective hypoglycemic agent. Metabolism. 1978;27:1539–1556. doi: 10.1016/S0026-0495(78)80027-4. [DOI] [PubMed] [Google Scholar]
- 85.Kiorpes TC, Hoerr D, Ho W, Weaner LE, Inman MG, Tutwiler GF. Identification of 2-tetradecylglycidyl coenzyme A as the active form of methyl 2-tetradecylglycidate (methyl palmoxirate) and its characterization as an irreversible, active site-directed inhibitor of carnitine palmitoyltransferase A in isolated rat liver mitochondria. J Biol Chem. 1884;259:9750–9755. [PubMed] [Google Scholar]
- 86.Selby PL, Sherratt HS. Substituted 2-oxiranecarboxylic acids: a new group of candidate hypoglycaemic drugs. Trends Pharmacol Sci. 1989;10:495–500. doi: 10.1016/0165-6147(89)90049-7. [DOI] [PubMed] [Google Scholar]
- 87.Bone AJ, Sherratt HS, Turnbull DM, Osmundsen H. Increased activity of peroxisomal beta-oxidation in rat liver caused by ethyl 2(5(4-chlorophenyl)pentyl)-oxiran-2-carboxylate: an inhibitor of mitochondrial beta-oxidation. Biochem Biophys Res Commun. 1982;104:708–712. doi: 10.1016/0006-291X(82)90694-5. [DOI] [PubMed] [Google Scholar]
- 88.Koundakjian PP, Turnbull DM, Bone AJ, Rogers MP, Younan SI, Sherratt HS. Metabolic changes in fed rats caused by chronic administration of ethyl 2[5(4-chlorophenyl)pentyl]oxirane-2-carboxylate, a new hypoglycaemic compound. Biochem Pharmacol. 1984;33:465–473. doi: 10.1016/0006-2952(84)90242-9. [DOI] [PubMed] [Google Scholar]
- 89.Wolf HP, Engel DW. Decrease of fatty acid oxidation, ketogenesis and gluconeogenesis in isolated perfused rat liver by phenylalkyl oxirane carboxylate (B 807–27) due to inhibition of CPT I (EC 2.3.1.21) Eur J Biochem. 1985;146:359–363. doi: 10.1111/j.1432-1033.1985.tb08661.x. [DOI] [PubMed] [Google Scholar]
- 90.Weis BC, Esser V, Foster DW, McGarry JD. Rat heart expresses two forms of mitochondrial carnitine palmitoyltransferase I. The minor component is identical to the liver enzyme. J Biol Chem. 1994;269:18712–18715. [PubMed] [Google Scholar]
- 91.Weis BC, Cowan AT, Brown N, Foster DW, McGarry JD. Use of a selective inhibitor of liver carnitine palmitoyltransferase I (CPT I) allows quantification of its contribution to total CPT I activity in rat heart. Evidence that the dominant cardiac CPT I isoform is identical to the skeletal muscle enzyme. J Biol Chem. 1994;269:26443–26448. [PubMed] [Google Scholar]
- 92.McGarry JD, Brown NF. The mitochondrial carnitine palmitoyltransferase system. From concept to molecular analysis. Eur J Biochem. 1997;244:1–14. doi: 10.1111/j.1432-1033.1997.00001.x. [DOI] [PubMed] [Google Scholar]
- 93.Merrill CL, Ni H, Yoon LW, Tirmenstein MA, Narayanan P, Benavides GR, Easton MJ, Creech DR, Hu CX, McFarland DC, Hahn LM, Thomas HC, Morgan KT. Etomoxir-induced oxidative stress in HepG2 cells detected by differential gene expression is confirmed biochemically. Toxicol Sci. 2002;68:93–101. doi: 10.1093/toxsci/68.1.93. [DOI] [PubMed] [Google Scholar]
- 94.Vickers AE, Bentley P, Fisher RL. Consequences of mitochondrial injury induced by pharmaceutical fatty acid oxidation inhibitors is characterized in human and rat liver slices. Toxicol In Vitro. 2006;20:1173–1182. doi: 10.1016/j.tiv.2006.01.021. [DOI] [PubMed] [Google Scholar]
- 95.Randle PJ. Endocrine control of metabolism. Annu Rev Physiol. 1963;25:291–324. doi: 10.1146/annurev.ph.25.030163.001451. [DOI] [PubMed] [Google Scholar]
- 96.Randle PJ. Regulatory interactions between lipids and carbohydrates: the glucose fatty acid cycle after 35 years. Diabetes Metab Rev. 1998;14:263–283. doi: 10.1002/(SICI)1099-0895(199812)14:4<263::AID-DMR233>3.0.CO;2-C. [DOI] [PubMed] [Google Scholar]
- 97.Ashrafian H, Horowitz JD, Frenneaux MP. Perhexiline. Cardiovasc Drug Rev. 2007;25:76–97. doi: 10.1111/j.1527-3466.2007.00006.x. [DOI] [PubMed] [Google Scholar]
- 98.Inglis S, Stewart S. Metabolic therapeutics in angina pectoris: history revisited with perhexiline. Eur J Cardiovasc Nurs. 2006;5:175–184. doi: 10.1016/j.ejcnurse.2006.01.001. [DOI] [PubMed] [Google Scholar]
- 99.Stephens TW, Higgins AJ, Cook GA, Harris RA. Two mechanisms produce tissue-specific inhibition of fatty acid oxidation by oxfenicine. Biochem J. 1985;227:651–660. doi: 10.1042/bj2270651. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Dai Y, Wolfgang MJ, Cha SH, Lane MD. Localization and effect of ectopic expression of CPT1c in CNS feeding centers. Biochem Biophys Res Commun. 2007;359:469–474. doi: 10.1016/j.bbrc.2007.05.161. [DOI] [PubMed] [Google Scholar]
- 101.Aja S, Landree LE, Kleman AM, Medghalchi SM, Vadlamudi A, McFadden JM, Aplasca A, Hyun J, Plummer E, Daniels K, Kemm M, Townsend CA, Thupari JN, Kuhajda FP, Moran TH, Ronnett GV. Pharmacological stimulation of brain carnitine palmitoyl-transferase-1 decreases food intake and body weight. Am J Physiol Regul Integr Comp Physiol. 2008;294:R352–R361. doi: 10.1152/ajpregu.00862.2006. [DOI] [PubMed] [Google Scholar]
- 102.López M, Lage R, Saha AK, Pérez-Tilve D, Vázquez MJ, Varela L, Sangiao-Alvarellos S, Tovar S, Raghay K, Rodríguez-Cuenca S, Deoliveira RM, Castañeda T, Datta R, Dong JZ, Culler M, Sleeman MW, Alvarez CV, Gallego R, Lelliott CJ, Carling D, Tschöp MH, Diéguez C, Vidal-Puig A. Hypothalamic fatty acid metabolism mediates the orexigenic action of ghrelin. Cell Metab. 2008;7:389–399. doi: 10.1016/j.cmet.2008.03.006. [DOI] [PubMed] [Google Scholar]
- 103.Vladutiu GD, Simmons Z, Isackson PJ, Tarnopolsky M, Peltier WL, Barboi AC, Sripathi N, Wortmann RL, Phillips PS. Genetic risk factors associated with lipid-lowering drug-induced myopathies. Muscle Nerve. 2006;34:153–162. doi: 10.1002/mus.20567. [DOI] [PubMed] [Google Scholar]
- 104.Wilke RA, Lin DW, Roden DM, Watkins PB, Flockhart D, Zineh I, Giacomini KM, Krauss RM. Identifying genetic risk factors for serious adverse drug reactions: current progress and challenges. Nat Rev Drug Discov. 2007;6:904–916. doi: 10.1038/nrd2423. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Antons KA, Williams CD, Baker SK, Phillips PS. Clinical perspectives of statin-induced rhabdomyolysis. Am J Med. 2006;119:400–409. doi: 10.1016/j.amjmed.2006.02.007. [DOI] [PubMed] [Google Scholar]
- 106.Bergstrom JD, Reitz RC. Studies on carnitine palmitoyl transferase: the similar nature of CPTi (inner form) and CPTo (outer form) Arch Biochem Biophys. 1980;204:71–79. doi: 10.1016/0003-9861(80)90008-9. [DOI] [PubMed] [Google Scholar]
- 107.Miyazawa S, Ozasa H, Osumi T, Hashimoto T. Purification and properties of carnitine octanoyltransferase and carnitine palmitoyltransferase from rat liver. J Biochem. 1983;94:529–542. doi: 10.1093/oxfordjournals.jbchem.a134384. [DOI] [PubMed] [Google Scholar]
- 108.Ramsay RR, Derrick JP, Friend AS, Tubbs PK. Purification and properties of the soluble carnitine palmitoyltransferase from bovine liver mitochondria. Biochem J. 1987;244:271–278. doi: 10.1042/bj2440271. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Finocchiaro G, Colombo I, DiDonato S. Purification, characterization and partial amino acid sequences of carnitine palmitoyl-transferase from human liver. FEBS Lett. 1990;274:163–166. doi: 10.1016/0014-5793(90)81354-Q. [DOI] [PubMed] [Google Scholar]
- 110.Woeltje KF, Kuwajima M, Foster DW, McGarry JD. Characterization of the mitochondrial carnitine palmitoyltransferase enzyme system. II. Use of detergents and antibodies. J Biol Chem. 1987;262:9822–9827. [PubMed] [Google Scholar]
- 111.Dobbins RL, Szczepaniak LS, Bentley B, Esser V, Myhill J, McGarry JD. Prolonged inhibition of muscle carnitine palmitoyltransferase-1 promotes intramyocellular lipid accumulation and insulin resistance in rats. Diabetes. 2001;50:123–130. doi: 10.2337/diabetes.50.1.123. [DOI] [PubMed] [Google Scholar]