Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2007 Nov 1.
Published in final edited form as: J Mol Cell Cardiol. 2006 Sep 26;41(5):762–773. doi: 10.1016/j.yjmcc.2006.08.006

Thioredoxin and Ventricular Remodeling

Tetsuro Ago 1, Junichi Sadoshima 1
PMCID: PMC1852508  NIHMSID: NIHMS13938  PMID: 17007870

Abstract

Increasing bodies of evidence indicate that reactive oxygen species (ROS) produced by mitochondria and other sources play an essential role in mediating ventricular remodeling after myocardial infarction and the development of heart failure. Antioxidants scavenge ROS, thereby maintaining the reduced environment of cells and inhibiting ventricular remodeling in the heart. Thioredoxin not only functions as a major antioxidant in the heart but also interacts with important signaling molecules and transcription factors, thereby modulating various cellular functions. The activity of thioredoxin is regulated by a variety of mechanisms, such as transcription, localization, protein-protein interaction, and post-translational modification. In this review, we will summarize the cardiac effects of thioredoxin and the mechanisms by which thioredoxin mediates inhibition of ventricular remodeling.

Keywords: thioredoxin, ventricular remodeling, cardiac hypertrophy, heart failure

1. Thioredoxin

Thioredoxin (Trx) is a major antioxidant in cells along with the tripeptide glutathione (GSH, Glu-Cys-Gly) [13]. These molecules maintain the intracellular space to be in reduced states [1,4]. Trx contains dithiol within its conserved active site (-Cys-Gly-Pro-Cys-; Cys at 32 and at 35 in Trx1) and reduces thiol-groups in various redox-sensitive proteins, while Trx itself is oxidized to form disulfide [1,5]. To work properly, Trx requires thioredoxin reductase (TrxR) and NADPH: TrxR reduces the active site of oxidized Trx using NADPH, thereby regenerating reduced Trx [2]. Therefore, the expression and activity of TrxR and the abundance of NADPH may affect Trx function. “The Trx system”, comprised of these three molecules, is conserved from E.coli to mammals, and is ubiquitously present in various tissues [6,7], which implicates the importance of Trx in cellular functions. In mammals, there are at least 3 members in the Trx family [3]: Trx1, Trx2, and Sp-trx (sperm trx, also designated as p32TrxL) [6,8,9]. In the broad sense, the Trx system may include thioredoxin peroxidase (peroxiredoxin): Trx reduces peroxiredoxin, which then catalyzes H2O2 to produce water [3].

1.1. Trx1

Among the three Trx family proteins, Trx1 has been most intensively examined thus far. Mammalian Trx1 is a small 12kDa protein, which was originally identified as adult T cell leukemia-derived factor produced by human T cell lymphotrophic virus-I-transformed T cells [10]. Trx1 activity is regulated by its expression level, localization (nucleus, cytosol, or extracellular space), interaction with other molecules, and post-translational modification (Figures 1 and 2).

Figure 1. Anti-hypertrophic and anti-apoptotic effects of Trx1.

Figure 1

Trx1 is normally localized in the cytosol, but translocates to the nucleus under stress conditions through direct interaction with importin (imp) or TXNIP. In the nucleus, Trx1 regulates the activity of various transcriptional factors. In the cytosol, Trx1 exerts anti-hypertrophic and anti-apoptotic effects by interacting with important signaling molecules, such as ASK-1, Ras, PTEN/Akt, and NFκB. Trx1 is also secreted to the extracellular space under stress conditions, thereby protecting neighboring cells.

Figure 2. Post-translational modifications of Trx1.

Figure 2

Trx1 has five modifiable cysteines (closed circles). Two of them, namely Cys-32 and Cys-35, in the catalytic site are involved in the enzymatic activity of Trx1. Under oxidative stresses, Trx1 forms disulfide bonds (S-S) between Cys-32 and Cys-35 and/or between Cys-62 and Cys-69, or a mixed disulfide bond with glutathione, called glutathionylation (S-SG). The activity of Trx1 or the accessibility of TrxR1 to Trx1 is inhibited by these oxidative modifications. On the other hand, Trx1 can be modified by NO and peroxynitrite. Trx1 is nitrosylated at Cys-69 (S-NO), which leads to the enhanced activity of Trx1. Trx1 has Tyr-49, which is nitrated by peroxynitrite. Nitration of Tyr-49 irreversibly suppresses the activity of Trx1.

1.1.1. Regulation of Trx1 expression

The promoter region of the trx1 gene contains a series of stress-responsive elements [11]. Trx1 transcript is induced by various stimuli and/or stresses, such as TNF-α, H2O2, UV, and heat shock [1,1214]. In addition, estrogen, prostaglandins E1, and cAMP are reported to induce mRNA, protein, and secretion of Trx1 [11,15]. Geranylgeranylacetone, a natural plant constituent and an anti-ulcer drug, is a molecule known to induce Trx1 expression [11,12,16]. In in vivo mouse models, pressure-overload and ischemia and reperfusion (I/R) induces Trx1 in the heart [17,18]. Among transcription factors, heat shock factor 2 (HSF2) is reported to induce Trx1 expression [19].

1.1.2. Localization of Trx1

Trx1 is normally present in the cytosol, but translocates into the nucleus under stress conditions [13], despite having no known nuclear localization sequence. A basic region, including Lys-81, 82, and 85, may interact with importin, a critical component of the nuclear transport system, thereby participating in nuclear localization of Trx1 [20]. Alternatively, the nuclear localization of Trx1 may be mediated by Thioredoxin1-interacting protein (TXNIP), which interacts with both Trx1 and importin α1 [21,22] (see section 6.2 and Figure 1). In the nucleus, Trx1 regulates the activity of various transcription factors, such as nuclear factor κB (NFκB), activator protein-1 (AP-1), and p53, in a redox-sensitive manner [23, 24], although the activation pattern of transcription factors by Trx1 may be cell type-dependent (see section 6.4 and Figure 1). Furthermore, Trx1 can be secreted to the extracellular space under stress conditions [11], and shows cytoprotective [23], anti-apoptotic [24], anti-inflammatory activity [25], as well as a growth-promoting effect as an autocrine/paracrine factor [26] (Figure 1). However, it is not well understood as to how Trx1 is secreted, and how extracellular Trx1 transmits its signal into the intracellular space. The extracellular function of Trx1 may be unique among various antioxidants, allowing Trx1 to alert neighboring cells and possibly prevent the expansion of the disaster [27].

1.1.3. Regulated activity of Trx1 by its interacting protein

The activity of Trx1 is negatively regulated by TXNIP at least at cellular levels [28,29] (see section 6.2 and Figure 1). Although TXNIP directly binds to a reduced form of Trx1, presumably at its catalytic center [28,29], the exact molecular mechanism by which TXNIP inhibits the protein–disulfide oxidoreductase activity of Trx1 remains to be elucidated. TXNIP existing in the cytoplasm also suppresses nuclear translocation of Trx1, thereby inhibiting nuclear effects of Trx1 in some cell types [30], although this observation contradicts the view that TXNIP exists primarily in the nucleus [21].

1.1.4. Post-translational modification of Trx1

Redox-based modifications are reported to regulate Trx1 activity (Figure 2). Besides the two cysteine residues in the catalytic site, Trx1 harbors three other cysteines, Cys-62, Cys-69, and Cys-73, which can be modified by ROS and nitric oxide (NO). It is suggested that Cys-62 and Cys-69 form another disulfide bond in the presence of oxidative stress, which may attenuate the accessibility of TrxR1 to Trx1, thereby suppressing Trx1 activity [31]. In addition, under oxidized conditions, Cys-73 reacts with glutathione to form a mixed disulfide, i.e. glutathionylation (S-SG), resulting in inhibition of the enzymatic activity [32]. In contrast, Cys-69 can be nitrosylated by NO, which may be required for scavenging ROS and for preserving its redox regulatory activity [33] (Figure 2). For example, in hearts subjected to ischemia/reperfusion, pretreatment of human Trx1 with s-nitrosoglutathione, which leads to nitro sylation of Trx1 at Cys-69, enhances the cardioprotective effect of Trx1 applied through intraperitoneal injection [24]. In contrast, E. coli Trx, which lacks Cys-69, does not possess this cardioprotective role [24]. Very recently, it has been reported that Trx1 undergoes irreversible nitrosative modification (Tyr-NO2) by peroxynitrite at Tyr-49 in the region critical for protein folding. In contrast to thiol-nitrosylation at Cys-69, the tyrosine nitration inhibits the activity of Trx1 [34]. Thus, a balance of ROS and NO in cells may determine the activity of Trx1.

1.2. Trx2

Trx2 is targeted to the mitochondria by its N-terminal mitochondria localization sequence, which is cleaved upon mitochondrial import. Trx2 contains the catalytic active site, but lacks the other modifiable cysteines found in Trx1 [3,35]. The mitochondrial Trx2 system, comprising Trx2, TrxR2, NADPH and peroxiredoxin-3 [36], is ubiquitously expressed, but very highly expressed in the metabolically active tissues, such as the heart, the brain and the liver. Mitochondria convert 1–2% of the oxygen molecules consumed into superoxide anion [37]. The superoxide production occurs in the electron transport chain, especially at complex I (NADH dehydrogenase) and complex III (ubiquinone-cytochrome c reductase). In addition, mitochondria are very vulnerable to ROS. Since the heart is one of the oxygen-consuming organs and its energy is absolutely dependent on ATP made in mitochondria [38], Trx2 appears to play an important role in protecting mitochondria from damages by ROS in the heart along with other mitochondrially localized antioxidant proteins, such as mitochondrial superoxide dismutase (SOD) [39], glutaredoxin-2 [40], and catalase [41].

1.3. KO mice in the thioredoxin system

1.3.1. Trx1 and TrxR1

While heterozygous deletion of Trx1 in mice causes no obvious phenotype, the homozygous mutants die shortly after implantation [42]. Trx1 expression seems to be essential for early differentiation and morphogenesis of the mouse embryo [42]. Although cardiac specific deletion of Trx1 has yet to be reported, mice with cardiac specific overexpression of mutant Trx1 that suppresses endogenous Trx1 activity display hypertrophy [17] (see section 3.1).

Ubiquitous inactivation of TrxR1 causes early embryonic lethality between 9.5 and 10.5 days in mice [43]. In contrast, mice with heart-specific inactivation of TrxR1 develop normally and appear healthy, suggesting that TrxR1 is dispensable at least for cardiac development [43]. By extension, if TrxR1 is only the reductase responsible for Trx1, Trx1 may be dispensable for development of the heart.

1.3.2. Trx2 and TrxR2

Cells deficient in Trx2 display increased cellular ROS and apoptosis [44]. In addition, The Trx2−/− homozygous mutant embryos display increased apoptosis at 10.5 days and are not present by 12.5 days [45]. The timing of the embryonic lethality is consistent with the maturation of the mitochondria, suggesting that Trx2 is needed for mitochondrial function.

TrxR2 knockout mice display embryonic death at embryonic day 13 [46]. Cardiac specific ablation of TrxR2 results in fatal dilated cardiomyopathy and death shortly after birth, suggesting that removal of toxic ROS species from mitochondria by TrxR2 (and also Trx2) is critical for cardiac function, which is absolutely dependent upon the maintenance of mitochondrial integrity [46].

2. “Ventricular remodeling” and ROS

Postnatal growth of the heart is primarily due to non-proliferative cardiac myocyte enlargement, termed hypertrophy, in response to mechanical or hormonal factors. Cardiac hypertrophy is also induced by many pathological insults, and is often accompanied by functional deterioration as well as histopathological changes, including apoptosis and fibrosis, a step called “ventricular remodeling” [47]. Besides structural remodeling, an altered utilization of the metabolic fuels from fatty acid to glucose, termed “metabolic remodeling”, occurs during hypertrophy and heart failure [48,49] (see section 4). Although ventricular remodeling may be initially an adaptive and beneficial physiological event for the maintenance of cardiac function, sustained remodeling processes often lead to pathological outcomes, i.e. heart failure [50].

Although ROS at low concentrations could mediate physiological cellular functions, high concentrations of ROS are detrimental. ROS may be involved in both steps of ventricular remodeling, i.e. developments of hypertrophy and transition from compensated hypertrophy to heart failure [5155]. ROS are generated by many hypertrophic stimuli, such as Gαq/Gα11-coupled receptor agonists (e.g., phenylephrine, angiotensin II) [56,57] and mechanical forces [58]. These ROS are generated by mitochondrial electron transport leakage [59,60], NAD(P)H oxidases [6163], xanthine oxidase [64], and uncoupled nitric oxide synthase (NOS) [65,66].

3. Roles of Trx1 in ventricular remodeling

Since increased production of ROS plays an important role in development of ventricular remodeling, antioxidants, such as glutathione, thioredoxin, SOD [67], catalase, glutaredoxin, and peroxiredoxin [36], are expected to play counteracting and thus protective roles. Interestingly, although a low dose of ROS induces Trx1 expression [14], a high dose of ROS inactivates Trx1 through post-translational modifications caused by imbalance between ROS and NO (see Section 1.1.4). Thus, the reactivity of Trx1 against different levels of ROS may differentially affect cellular responses of cardiac myocytes, i.e. hypertrophy or apoptosis.

3.1. Cardiac hypertrophy and Trx1

Since a low dose of ROS promotes cell growth [2,38,62], antioxidants, such as Trx1, may attenuate cell growth. However, Trx1 has the potential to promote cell growth in some cell types, such as cancer cells [1,12,68]. Thus, these conflicting effects of Trx1 raise the following question, “What is the net effect of Trx1 on cardiac hypertrophy?”

In order to address this question, we made transgenic mice with cardiac specific overexpression of Trx1 (Tg-Trx1) and catalytically defective Trx1 harboring the double substitution of Cys32/35Ser (Tg-Trx1-DN) [17]. Interestingly, suppression of endogenous Trx1 by overexpression of Trx1-DN induces cardiac hypertrophy at baseline and exacerbates hypertrophic response induced by thoracic aortic constriction (TAC). On the other hand, baseline cardiac hypertrophy was never observed in Tg-Trx1 mice, while cardiac hypertrophy induced by TAC is significantly suppressed in Tg-Trx1 mice [17]. These findings indicate that Trx1 suppresses pathological cardiac hypertrophy. It is not clear why cardiac myocyte apoptosis is not prominent in Tg-Trx1-DN mice, even though ROS are increased in the mouse hearts. The level of ROS produced by suppression of Trx1 may not be sufficient to induce apoptosis. Alternatively, ROS may be produced in a subcellular compartment which couples to hypertrophy but not apoptosis. A possible mechanism underlying the Trx1-mediated anti-hypertrophic effect could be inhibition of the Ras-MAPK signaling pathway [17]. Consistent with our findings, Kuster et al have shown that Trx1 inhibits Ras activation induced by α-adrenergic receptor stimulation, thereby suppressing hypertrophy in adult rat ventricular myocytes [69]. Ras has redox-sensitive cysteine residues, and during α-adrenergic receptor-stimulated hypertrophy, ROS, probably produced by NAD(P)H oxidase or mitochondria, oxidize the thiols in the cysteines, leading to enhanced activation of Ras [63]. By contrast, Trx1 directly reduces the thiols in Ras, thereby inhibiting Ras [69] (Figure 1). It should be noted, however, that this may not be the sole mechanism by which Trx1 inhibits cardiac hypertrophy, because Trx1 modulates various protein-protein interactions and signaling pathways.

3.2. Ischemia/reperfusion injury, preconditioning and Trx1

Preconditioning is the phenomenon that sublethal stress increases the tolerance of myocardium against a subsequent prolonged ischemic insult [70,71]. Preconditioning is one of the most powerful mechanisms of protection of heart cells from injury and death induced by I/R. Both ROS and Trx1 may play indispensable roles in mediating preconditioning. Molecules released from cells subjected to I/R, including adenosine, angiotensin II, and nitric oxide [7072], often activate protein kinase C, a major mediator of preconditioning, which is involved in ROS production during ischemic preconditioning [7376]. On the other hand, cardioprotective effects of preconditioning are abolished when the heart is pre-perfused with N-acetyl cysteine, a scavenger of ROS [77]. Thus, ROS production is required for initiation of preconditioning. Turoczi et al reported, using ex vivo working rat hearts, that Trx1 was upregulated by ischemic preconditioning [18]. In addition, overexpression of Trx1 improved post-ischemic ventricular recovery and reduced myocardial infarct size in mouse hearts, implicating a protective role of Trx1 in ischemic myocardium [18]. Furthermore, Malik et al showed that inhibition of Trx1 by RNA interference attenuates the cardioprotective effect of Trx1 during preconditioning, where nuclear translocation of Trx1 and subsequent activation of NFκB may be critical for cardioprotection [78] (Figure 1). Das and colleagues propose that Trx1 plays a critical role in mediating the preconditioning of myocardial ischemia [72,77].

Plasma Trx1 levels are elevated in patients with acute myocardial infarction [79]. The elevated levels of Trx1 may be protective under ischemic conditions, because exogenously applied human Trx1 was incorporated into myocardial tissue and reduced both infarct size induced by I/R [24] and occurrence of arrhythmia [80].

3.3. Heart failure and Trx1

Chronic heart failure is a complex neuro-hormonal and inflammatory syndrome [81]. Oxidative stress is enhanced during chronic heart failure [62], where production of ROS is enhanced or the activity of antioxidants is suppressed. ROS not only induce apoptosis of cardiac myocytes [82] but also stimulate fibrosis [23]. ROS also elicit reduced cardiac pump function [83] through suppressed activity of ion channels, such as L-type calcium channels on the sarcolemma [84], and of the sarcoplasmic reticulum Ca2+ ATPase, SERCA2 [85]. During hypertrophy and heart failure, ROS are produced primarily by mitochondria [59,62], NAD(P)H oxidase [61,62] and uncoupled eNOS [66]. High amounts of ROS further damage mitochondria, through ROS-induced ROS production, thereby forming a vicious cycle of ROS and mitochondrial dysfunction [86].

Although the exact tissue level of Trx1 during heart failure in humans has yet to be elucidated, serum concentrations of Trx1 are known to be increased in patients with heart failure and correlate with the severity of the disease [87,88]. The upregulation of Trx1 is assumed to counteract increased oxidative stress in cardiac myocytes and myocardium. Circulating Trx1 suppresses inflammation by blocking chemotaxis of neutrophils [25,89] and death of cardiac myocytes [24] (Figure 1). However, there has been no direct evidence so far to show that Trx1 prevents the occurrence and progression of heart failure. We have recently found, using a mouse model of dilated cardiomyopathy in which mammalian sterile 20 like kinase 1 (Mst1) is overexpressed specifically in the heart [90], that overexpression of Trx1 prevents progression of heart failure in the Mst1 mice (Ago and Sadoshima, unpublished data).

3.3.1. Apoptosis and Trx1

Apoptosis is one of the important components of heart failure. Trx1 exerts an anti-apoptotic effect through its antioxidant effect [33] and interaction with critical proteins that regulate apoptosis, such as apoptosis signal-regulating kinase 1 (ASK-1) [9194] and Akt [95,96]. The reduced form of Trx1 inactivates ASK-1, an upstream kinase that leads to activation of p38 and JNK, critical mediators of apoptosis in cardiac myocytes, thereby inhibiting apoptosis [91]. Trx1 appears to activate Akt [95], which directly phosphorylates and inactivates Bad, a pro-apoptotic protein of the Bcl-2 family [97], and procaspase-9, a critical component of apoptosomes [98]. Trx1 enhances the DNA binding activity of NFκB, which is anti-apoptotic under some conditions, such as I/R, by inducing anti-apoptotic genes [99]. The anti-apoptotic mechanisms involving these molecules induced by Trx1 will also be discussed in section 6 (Figure 1).

3.3.2. Cardiac fibrosis and Trx1

The extracellular matrix, produced by cardiac fibroblasts, normally maintains the myocardial architecture. However, increased synthesis of collagen causes fibrosis. Matrix metalloproteinase (MMPs) and tissue inhibitors of matrix metalloproteinases (TIMPs) mainly regulate collagen deposition and degradation. Dysregulation of the extracellular matrix by these molecules contributes to abnormal cardiac muscle fibrosis, leading to progression of heart failure. It is reported that ROS activate expression of MMPs, such as MMP-2 and MMP-9 [100], whereas ROS decrease TIMP expression and fibrillar collagen synthesis in cardiac fibroblasts [101,102]. Furthermore, ROS directly activate MMPs post-translationally [102]. In contrast, inhibition of ROS derived from NAD(P)H oxidase [103] and mitochondria [36] leads to reduced myocardial remodeling in animal models [102]. Trx1 prevents pro-inflammatory cytokine- or bleomycin-induced lung fibrosis, which is mediated by ROS [104]. Trx1 suppresses progression of cardiac fibrosis in a mouse model of heart failure, although the underlying molecular mechanism has yet to be elucidated (our unpublished data).

3.4. Aging heart and Trx1

One of the major hypotheses regarding the molecular mechanism of aging is that oxidative stress underlies aging processes [4]. Thus, protective mechanisms against oxidative stress may be important in counteracting the aging process [4,105]. Systemic overexpression of Trx1 in transgenic mice resulted in extended median and maximum life spans compared with wild-type mice [106,107]. Interestingly, although enhanced telomerase activity was found in the Trx1 transgenic mice, proliferative changes, especially cancer, do not frequently occur [106]. In the heart, Trx1 is expected to attenuate structural and functional changes with aging. Indeed, Trx1 appears to attenuate age-induced cardiac hypertrophy in mice (Ago and Sadoshima, unpublished data). It has been shown that expression of TrxR2 is severely attenuated by aging, a condition which is normalized by caloric restriction, a factor for longevity [108].

Trx1 also serves as an electron donor for methionine sulfoxide reductases (Msr) as well as peroxiredoxins [109,110]. Msr, expressed in all human tissues, including the heart, catalyzes the thioredoxin-dependent reduction of methionine sulfoxide residues back to methionine residues [111]. The importance of Msr is highlighted by the fact that aging is associated with a loss of Msr activity in a number of animal tissues, and that C57BL/J6 mice lacking Msr display a decreased maximum life span [112], whereas overexpression of Msr leads to a dramatic increase in the maximum life span [113]. Thus, a possible mechanism by which Trx1 antagonizes aging may be due to keeping Msr in an active state.

4. Metabolic remodeling during hypertrophy and heart failure

Cardiac muscle needs a high turnover of ATP made by mitochondria in order to contract efficiently. Under aerobic conditions, fatty acid oxidation accounts for up to 90% of the total acetyl-CoA, the substrate of the TCA cycle [114]. However, during cardiac hypertrophy and heart failure, the metabolic process to supply acetyl-CoA in the heart shifts from fatty acid oxidation to glucose-mediated metabolism, termed metabolic remodeling [48], which is accompanied by a degree of cellular oxidative stress [115]. Alteration of gene expression responsible for metabolic remodeling during cardiac hypertrophy and heart failure is in part mediated by the fatty acid-activated peroxisome proliferator-activated receptors (PPARs) and PPARγ coactivator-1α (PGC-1α), major regulators of mitochondrial biogenesis and respiration [116,117]. Expression of these molecules is extensively suppressed during hypertrophy and heart failure [49,117119]. Cardiac specific deletion of either PPARs [120122] or PGC-1α [123] leads to decreased fatty acid oxidation and subsequent cardiac dysfunction, suggesting that PPARs and PGC-1α primarily regulate cardiac function and prevent the progression of heart failure [124]. We have recently found, using cDNA microarray analyses, that expression of PGC-1α and the activity of mitochondrial enzymes are significantly upregulated by Trx1 overexpression in the heart [125]. This finding suggests that Trx1 may play a protective role against hypertrophy and heart failure by attenuating metabolic remodeling through an enhancement of PGC-1α activity (Figure 1).

5. Channel remodeling during hypertrophy and heart failure

Along with structural and metabolic remodeling, “channel remodeling” also occurs during hypertrophy and heart failure, and leads to contractile dysfunction and arrhythmias characterized by downregulation of K+ channels that control action potential morphology [126,127]. Li et al suggested that Trx1 upregulates expression of cardiac ion channels responsible for transient outward current, and reverses channel remodeling through its redox activity [128]. Thus, Trx1 may also work protectively against “channel remodeling”.

6. Trx1-interacting proteins involved in cardiac remodeling

Trx1 not only reduces redox-sensitive molecules but also interacts with various proteins, such as intracellular signaling molecules and transcription factors (Figure 1). The interaction is often dependent on the redox state and localization of Trx1. Through protein-protein interaction, Trx1 alters the enzymatic activity or the subcellular localization of the partner protein, thereby affecting various cellular functions.

6.1. ASK-1

ASK-1, a MAPKKK [129], is one of the most prominent molecules that are regulated by Trx1. Increasing lines of evidence suggest that ASK-1 plays an important role in mediating both cardiac hypertrophy and apoptosis [94]. For example, G-protein–coupled receptor (GPCR) agonists such as angiotensin II, endothelin-1 and phenylephrine generate ROS and activate ASK-1, thereby leading to cardiac myocyte hypertrophy [130] (Figure 1). In ASK-1 knockout mice, angiotensin II-induced cardiac hypertrophy and remodeling, as well as p38 and JNK activation, are significantly attenuated [131]. Trx1 and ASK-1 physically interact with one another and this interaction is redox-dependent: oxidized Trx1 is unable to interact with ASK-1 [91]. Through this direct interaction, Trx1 suppresses the kinase activity of ASK-1, thereby inhibiting the apoptotic as well as hypertrophic function of ASK-1 [91,129] (Figure 1). In addition, Trx1 may promote ubiquitination and degradation of ASK-1, a function for which a single cysteine in the catalytic site of Trx1 is sufficient [92]. Thus, both anti-hypertrophic and anti-apoptotic actions of Trx1 may be mediated at least in part by the interaction with ASK-1. ASK-1 associates with troponin T and phosphorylates it, thereby reducing contractility in cardiac myocytes [132], which may also be attenuated by Trx1. Recently, it has been shown that Trx2 also interacts with ASK-1 in the mitochondria, thereby inhibiting ASK1-induced apoptosis [93].

6.2. TXNIP/ VDUP1/TBP-2

TXNIP was originally identified as an upregulated protein in HL-60, human promyelocytic leukemia cells, treated with 1α, 25 dihydroxyvitamin D3, and was thus designated vitamin D3-upregulated protein 1 (VDUP1) [133]. Using a yeast two hybrid system, Nishiyama et al identified this same protein as a Trx1-binding protein, and originally named it TBP-2 [28]. TXNIP binds to reduced but not oxidized Trx1, suggesting that the catalytic center of Trx1 is important for this interaction [28]. Although it is thought that TXNIP attenuates Trx1 activity, at least as measured by insulin-reducing activity at in vitro and cellular levels [28,29], Trx1 activity is not enhanced in TXNIP knockout mice [134]. Expression of TXNIP in cardiac myocytes is rapidly suppressed by hypertrophic stimuli, such as mechanical strain, hydrogen peroxide, and pressure-overload [135,136]. In contrast, ectopic overexpression of TXNIP suppresses cellular proliferation in tumor cells [137] and, in cardiac myocytes, attenuates cardiac hypertrophy in response to mechanical strain, phenylephrine, or angiotensin II [136], and causes apoptotic cell death [135]. Thus, TXNIP serves as both a negative regulator of cardiac hypertrophy and a positive mediator of cell death. If TXNIP functions only to attenuate Trx1 activity, these findings may not be consistent with the reported Trx1 functions in the heart [17]. Rather, Trx1 and TXNIP may function in a cooperative fashion to suppress hypertrophy. Indeed, HSF, a transcription factor, upregulates both Trx1 [19] and TXNIP [138], and TXNIP may mediate the translocation of Trx1 through interaction with importin α1 into the nucleus, where Trx1 plays a critical role in regulating transcription [21] (Figure 1). However, it should be noted that TXNIP may inhibit nuclear translocation of Trx1 in some cell types [30]. It remains to be elucidated how Trx1 and TXNIP regulate their activities of one another during different disease states in the heart.

It is reported that the phenotype of the mutant mouse strain, HcB-19/Dem, which shows characteristics of familial combined hyperlipidemia, may be caused by a nonsense mutation in TXNIP [139]. Consistently, TXNIP knockout mice display impaired fatty acid oxidation and utilization by the TCA cycle, with resultant increased levels of fatty acid and reduced levels of glucose in plasma [134]. These findings suggest that TXNIP may regulate fatty acid metabolism and metabolic remodeling in the heart as well. As we mentioned, microarray analyses of Tg-Trx1 mice show that Trx1 increases PGC-1α expression [125], which may enhance fatty acid oxidation. Thus, it is possible that Trx1 and TXNIP may function cooperatively to maintain the normal metabolic state in the heart (Figure 1).

6.3. PTEN/PI3K/Akt pathway

Trx1 has been shown to interact with Phosphatase and Tensin homolog (PTEN), a protein phosphatase that attenuates PI3K/Akt activity [95]. PTEN has a critical cysteine at residue 212 in the C2 lipid-binding domain, which mediates interaction with Cys-32 in the catalytic site of Trx1 to form a disulfide bond. The interaction leads to steric hindrance of the catalytic site of PTEN [95]. Thus, the Trx1-PTEN interaction attenuates the phosphatase activity of PTEN, thereby activating PI3K and Akt (Figure 1). Since a signaling cascade that leads to Akt activation promotes growth and survival of many cell types, including cardiac myocytes, stimulation of Akt via suppression of PTEN may explain the anti-apoptotic effect of Trx1. There are some reports showing that ROS enhance Akt activity by directly inhibiting PTEN via oxidation [140]. Consistently, in ischemic preconditioning, oxidative stress, rather than Trx1 induction, may induce PTEN inactivation and further activation of Akt [141]. In these studies, however, a possibility remains that ROS-induced upregulation of Trx1 may mediate Akt activation.

Growing bodies of evidence suggest that the IGF-PI3K-Akt pathway mediates physiological rather than pathological hypertrophy [142]. Interestingly, akt1−/− mice develop an exacerbated form of cardiac hypertrophy in response to pressure-overload, suggesting an antagonizing effect of Akt against pathological hypertrophy, although physiological hypertrophy induced by swimming was significantly suppressed in these mice [143]. Thus, there is a possibility that Trx1-mediated activation of PI3K/Akt through PTEN inhibition may participate in physiological cardiac hypertrophy, while attenuating pathological hypertrophy as observed in Tg-Trx1 mice [17]. Alternatively, the anti-hypertrophic effect of Trx1 may be totally independent of the PTEN/PI3K/Akt pathway.

6.4. Transcriptional factors and their regulators

The activity of several transcription factors is altered by redox modification of specific cysteine residues. The DNA-binding activity of NFκB is markedly augmented by direct interaction with Trx1 in vitro [144]. However, Trx1 also inhibits NFκB activation: in contrast to the event in the nucleus, Trx1 blocks degradation of IκB and prevents nuclear translocation of NFκB [145] (Figure 1). The DNA-binding activity of AP-1 is modified by a DNA repair enzyme, redox factor 1 (Ref-1). Ref-1 activity is regulated by interaction with Trx1 [146]. Trx1 associates directly with Ref-1 in the nucleus only in its reduced form, which then potentiates AP-1 activity through direct interaction [147] (Figure 1). Because NFκB and AP-1 are well known transcription factors related to the development of cardiac hypertrophy [148,149], it is expected that Trx1 is involved in hypertrophic responses in the heart through these transcription factors. However, we have recently shown, using cDNA microarray analyses, that the activity of these transcription factors is not enhanced in the heart of Tg-Trx1 mice at baseline [125]. Thus, the functional outcome of transcription factor regulation by Trx1 may be different among various cell types and cell conditions (i.e. normal vs diseased states, such as I/R and heart failure). Using microarray analyses in conjunction with transcription factor binding site analyses, we have found that the transcriptional activity of CREB and nuclear respiratory factor 1 (NRF1) is enhanced in Tg-Trx1 mouse hearts [125] (Figure 1). Although it has been reported that Trx1 reduces a cysteine residue of NRF2, thereby enhancing the DNA binding activity of NRF2 [150], it remains unknown whether CREB and NRF1 also have similar redox sensitive cysteines regulated by Trx1.

7. Therapeutic implications

Trx1 may function as a negative regulator of hypertrophy and heart failure. Modulating the negative regulators of hypertrophy may become one of the important strategies for combating heart failure in the future [151]. Although some negative regulators of hypertrophy induce apoptosis [151,152], Trx1 suppresses both hypertrophy and apoptosis, which would be ideal when considered as a therapeutic agent for many heart diseases. In addition, Trx1 has advantages for use in clinical therapy: it is small-sized and structurally stable, and can function both intracellularly and extracellularly, which would allow administration of Trx1 directly into either the myocardium or coronary circulation. Indeed, in animal models, direct administration of Trx1 exhibited salutary effects on some heart disease states [12,24]. However, higher doses of Trx1 may be needed so far to reduce oxidative stress and inflammation [12]. In addition to direct infusion of Trx1, the administration of reagents that induce Trx1 expression is also worthy of attention. Geranylgeranylacetone is one such drug, which is already used as an anti-ulcer drug. Interestingly, resveratrol, a major component of red grapes, is known to induce Trx1 [153]. Finding the molecules that elicit Trx1 expression effectively and in a highly selective fashion may expand our choice for cardiac therapy. In summary, we believe that Trx1 may be one of the most promising molecules to be used for therapy for ischemic heart disease and congestive heart failure.

8. Concluding remarks

Trx1 has a wide variety of cellular functions, not only as an antioxidant but also as a regulator of diverse signaling pathways through direct interaction with intracellular signaling molecules and transcription factors. Through these functions, Trx1 prevents the development of ventricular remodeling and congestive heart failure. However, the molecular mechanisms by which Trx1 confers advantages to the heart are still not completely understood. Further investigations are needed in order for us to apply our knowledge regarding the potentially salutary effects of Trx1 to treatment for cardiovascular patients.

Acknowledgments

We thank Daniela Zablocki for critical reading of the manuscript. This work was supported in part by U.S. Public Health Service Grants HL 59139, HL67724, HL67727, HL69020, and HL 73048 and by the American Heart Association grant 0340123N. T.A. is supported in part by a Research Fellowship from the Uehara Memorial Foundation, Japan.

List of Abbreviations

AP-1

activator protein 1

ASK-1

apoptosis signal-regulating kinase 1

DN

dominant-negative

GPCR

G-protein–coupled receptor

HSF

heat shock factor

I/R

ischemia/reperfusion

MMP

matrix metalloproteinase

Msr

methionine sulfoxide reductases

NFκB

nuclear factor κB

NO

nitric oxide

NOS

nitric oxide synthase

PGC-1α

PPARγ coactivator-1α, PPAR, peroxisome proliferator-activated receptor

PTEN

Phosphatase and Tensin homolog

Ref-1

redox factor 1

ROS

reactive oxygen species

SOD

superoxide dismutase

TAC

thoracic aortic constriction

Tg-Trx1

transgenic mice with cardiac specific thioredoxin-overexpression

TIMP

tissue inhibitors of matrix metalloproteinase

Trx

thioredoxin

TrxR

thioredoxin reductase

TXNIP

thioredoxin-interacting protein

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Nakamura H, Nakamura K, Yodoi J. Redox regulation of cellular activation. Annu Rev Immunol. 1997;15:351–69. doi: 10.1146/annurev.immunol.15.1.351. [DOI] [PubMed] [Google Scholar]
  • 2.Nordberg J, Arner ES. Reactive oxygen species, antioxidants, and the mammalian thioredoxin system. Free Radic Biol Med. 2001;31:1287–312. doi: 10.1016/s0891-5849(01)00724-9. [DOI] [PubMed] [Google Scholar]
  • 3.Watson WH, Yang X, Choi YE, Jones DP, Kehrer JP. Thioredoxin and its role in toxicology. Toxicol Sci. 2004;78:3–14. doi: 10.1093/toxsci/kfh050. [DOI] [PubMed] [Google Scholar]
  • 4.Finkel T, Holbrook NJ. Oxidants, oxidative stress and the biology of ageing. Nature. 2000;408:239–47. doi: 10.1038/35041687. [DOI] [PubMed] [Google Scholar]
  • 5.Yamawaki H, Berk BC. Thioredoxin: a multifunctional antioxidant enzyme in kidney, heart and vessels. Curr Opin Nephrol Hypertens. 2005;14:149–53. doi: 10.1097/00041552-200503000-00010. [DOI] [PubMed] [Google Scholar]
  • 6.Powis G, Mustacich D, Coon A. The role of the redox protein thioredoxin in cell growth and cancer. Free Radic Biol Med. 2000;29:312–22. doi: 10.1016/s0891-5849(00)00313-0. [DOI] [PubMed] [Google Scholar]
  • 7.Holmgren A. Thioredoxin. Annu Rev Biochem. 1985;54:237–71. doi: 10.1146/annurev.bi.54.070185.001321. [DOI] [PubMed] [Google Scholar]
  • 8.Lee KK, Murakawa M, Takahashi S, Tsubuki S, Kawashima S, Sakamaki K, et al. Purification, molecular cloning, and characterization of TRP32, a novel thioredoxin-related mammalian protein of 32 kDa. J Biol Chem. 1998;273:19160–6. doi: 10.1074/jbc.273.30.19160. [DOI] [PubMed] [Google Scholar]
  • 9.Miranda-Vizuete A, Ljung J, Damdimopoulos AE, Gustafsson JA, Oko R, Pelto-Huikko M, et al. Characterization of Sptrx, a novel member of the thioredoxin family specifically expressed in human spermatozoa. J Biol Chem. 2001;276:31567–74. doi: 10.1074/jbc.M101760200. [DOI] [PubMed] [Google Scholar]
  • 10.Tagaya Y, Maeda Y, Mitsui A, Kondo N, Matsui H, Hamuro J, et al. ATL-derived factor (ADF), an IL-2 receptor/Tac inducer homologous to thioredoxin; possible involvement of dithiol-reduction in the IL-2 receptor induction. Embo J. 1989;8:757–64. doi: 10.1002/j.1460-2075.1989.tb03436.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Hirota K, Nakamura H, Masutani H, Yodoi J. Thioredoxin superfamily and thioredoxin-inducing agents. Ann N Y Acad Sci. 2002;957:189–99. doi: 10.1111/j.1749-6632.2002.tb02916.x. [DOI] [PubMed] [Google Scholar]
  • 12.Burke-Gaffney A, Callister ME, Nakamura H. Thioredoxin: friend or foe in human disease? Trends Pharmacol Sci. 2005;26:398–404. doi: 10.1016/j.tips.2005.06.005. [DOI] [PubMed] [Google Scholar]
  • 13.Tanaka T, Nishiyama Y, Okada K, Hirota K, Matsui M, Yodoi J, et al. Induction and nuclear translocation of thioredoxin by oxidative damage in the mouse kidney: independence of tubular necrosis and sulfhydryl depletion. Lab Invest. 1997;77:145–55. [PubMed] [Google Scholar]
  • 14.Haendeler J, Tischler V, Hoffmann J, Zeiher AM, Dimmeler S. Low doses of reactive oxygen species protect endothelial cells from apoptosis by increasing thioredoxin-1 expression. FEBS Lett. 2004;577:427–33. doi: 10.1016/j.febslet.2004.10.041. [DOI] [PubMed] [Google Scholar]
  • 15.Yamamoto M, Sato N, Tajima H, Furuke K, Ohira A, Honda Y, et al. Induction of human thioredoxin in cultured human retinal pigment epithelial cells through cyclic AMP-dependent pathway; involvement in the cytoprotective activity of prostaglandin E1. Exp Eye Res. 1997;65:645–52. doi: 10.1006/exer.1997.0370. [DOI] [PubMed] [Google Scholar]
  • 16.Hirota K, Nakamura H, Arai T, Ishii H, Bai J, Itoh T, et al. Geranylgeranylacetone enhances expression of thioredoxin and suppresses ethanol-induced cytotoxicity in cultured hepatocytes. Biochem Biophys Res Commun. 2000;275:825–30. doi: 10.1006/bbrc.2000.3392. [DOI] [PubMed] [Google Scholar]
  • 17.Yamamoto M, Yang G, Hong C, Liu J, Holle E, Yu X, et al. Inhibition of endogenous thioredoxin in the heart increases oxidative stress and cardiac hypertrophy. J Clin Invest. 2003;112:1395–406. doi: 10.1172/JCI17700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Turoczi T, Chang VW, Engelman RM, Maulik N, Ho YS, Das DK. Thioredoxin redox signaling in the ischemic heart: an insight with transgenic mice overexpressing Trx1. J Mol Cell Cardiol. 2003;35:695–704. doi: 10.1016/s0022-2828(03)00117-2. [DOI] [PubMed] [Google Scholar]
  • 19.Leppa S, Pirkkala L, Chow SC, Eriksson JE, Sistonen L. Thioredoxin is transcriptionally induced upon activation of heat shock factor 2. J Biol Chem. 1997;272:30400–4. doi: 10.1074/jbc.272.48.30400. [DOI] [PubMed] [Google Scholar]
  • 20.Haendeler J, Goy C, Dimmeler S. The anti-apoptotic function of physiological concentration of reactive oxygen species is dependent on the nuclear import of thioredoxin-1 in endothelial cells. Circulation. 2005;112:II-145. [Google Scholar]
  • 21.Nishinaka Y, Masutani H, Oka S, Matsuo Y, Yamaguchi Y, Nishio K, et al. Importin alpha1 (Rch1) mediates nuclear translocation of thioredoxin-binding protein-2/vitamin D(3)-up-regulated protein 1. J Biol Chem. 2004;279:37559–65. doi: 10.1074/jbc.M405473200. [DOI] [PubMed] [Google Scholar]
  • 22.Yoneda Y. Nucleocytoplasmic protein traffic and its significance to cell function. Genes Cells. 2000;5:777–87. doi: 10.1046/j.1365-2443.2000.00366.x. [DOI] [PubMed] [Google Scholar]
  • 23.Chen K, Chen J, Li D, Zhang X, Mehta JL. Angiotensin II regulation of collagen type I expression in cardiac fibroblasts: modulation by PPAR-gamma ligand pioglitazone. Hypertension. 2004;44:655–61. doi: 10.1161/01.HYP.0000144400.49062.6b. [DOI] [PubMed] [Google Scholar]
  • 24.Tao L, Gao E, Bryan NS, Qu Y, Liu HR, Hu A, et al. Cardioprotective effects of thioredoxin in myocardial ischemia and reperfusion: role of S-nitrosation [corrected] Proc Natl Acad Sci U S A. 2004;101:11471–6. doi: 10.1073/pnas.0402941101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Nakamura H, Herzenberg LA, Bai J, Araya S, Kondo N, Nishinaka Y, et al. Circulating thioredoxin suppresses lipopolysaccharide-induced neutrophil chemotaxis. Proc Natl Acad Sci U S A. 2001;98:15143–8. doi: 10.1073/pnas.191498798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Wakasugi N, Tagaya Y, Wakasugi H, Mitsui A, Maeda M, Yodoi J, et al. Adult T-cell leukemia-derived factor/thioredoxin, produced by both human T-lymphotropic virus type I- and Epstein-Barr virus-transformed lymphocytes, acts as an autocrine growth factor and synergizes with interleukin 1 and interleukin 2. Proc Natl Acad Sci U S A. 1990;87:8282–6. doi: 10.1073/pnas.87.21.8282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Sinclair DA. Toward a unified theory of caloric restriction and longevity regulation. Mech Ageing Dev. 2005;126:987–1002. doi: 10.1016/j.mad.2005.03.019. [DOI] [PubMed] [Google Scholar]
  • 28.Nishiyama A, Matsui M, Iwata S, Hirota K, Masutani H, Nakamura H, et al. Identification of thioredoxin-binding protein-2/vitamin D(3) up-regulated protein 1 as a negative regulator of thioredoxin function and expression. J Biol Chem. 1999;274:21645–50. doi: 10.1074/jbc.274.31.21645. [DOI] [PubMed] [Google Scholar]
  • 29.Junn E, Han SH, Im JY, Yang Y, Cho EW, Um HD, et al. Vitamin D3 up-regulated protein 1 mediates oxidative stress via suppressing the thioredoxin function. J Immunol. 2000;164:6287–95. doi: 10.4049/jimmunol.164.12.6287. [DOI] [PubMed] [Google Scholar]
  • 30.Schulze PC, De Keulenaer GW, Yoshioka J, Kassik KA, Lee RT. Vitamin D3-upregulated protein-1 (VDUP-1) regulates redox-dependent vascular smooth muscle cell proliferation through interaction with thioredoxin. Circ Res. 2002;91:689–95. doi: 10.1161/01.res.0000037982.55074.f6. [DOI] [PubMed] [Google Scholar]
  • 31.Watson WH, Pohl J, Montfort WR, Stuchlik O, Reed MS, Powis G, et al. Redox potential of human thioredoxin 1 and identification of a second dithiol/disulfide motif. J Biol Chem. 2003;278:33408–15. doi: 10.1074/jbc.M211107200. [DOI] [PubMed] [Google Scholar]
  • 32.Casagrande S, Bonetto V, Fratelli M, Gianazza E, Eberini I, Massignan T, et al. Glutathionylation of human thioredoxin: a possible crosstalk between the glutathione and thioredoxin systems. Proc Natl Acad Sci U S A. 2002;99:9745–9. doi: 10.1073/pnas.152168599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Haendeler J, Hoffmann J, Tischler V, Berk BC, Zeiher AM, Dimmeler S. Redox regulatory and anti-apoptotic functions of thioredoxin depend on S-nitrosylation at cysteine 69. Nat Cell Biol. 2002;4:743–9. doi: 10.1038/ncb851. [DOI] [PubMed] [Google Scholar]
  • 34.Tao L, Jiao X, Gao E, Lau W, Yuan Y, Lopez B, et al. Nitrative inactivation of thioredoxin-1 and its role in post-ischemic myocardial apoptosis. Circulation In press. doi: 10.1161/CIRCULATIONAHA.106.625061. [DOI] [PubMed] [Google Scholar]
  • 35.Spyrou G, Enmark E, Miranda-Vizuete A, Gustafsson J. Cloning and expression of a novel mammalian thioredoxin. J Biol Chem. 1997;272:2936–41. doi: 10.1074/jbc.272.5.2936. [DOI] [PubMed] [Google Scholar]
  • 36.Matsushima S, Ide T, Yamato M, Matsusaka H, Hattori F, Ikeuchi M, et al. Overexpression of mitochondrial peroxiredoxin-3 prevents left ventricular remodeling and failure after myocardial infarction in mice. Circulation. 2006;113:1779–86. doi: 10.1161/CIRCULATIONAHA.105.582239. [DOI] [PubMed] [Google Scholar]
  • 37.Boveris A, Chance B. The mitochondrial generation of hydrogen peroxide. General properties and effect of hyperbaric oxygen. Biochem J. 1973;134:707–16. doi: 10.1042/bj1340707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Giordano FJ. Oxygen, oxidative stress, hypoxia, and heart failure. J Clin Invest. 2005;115:500–8. doi: 10.1172/JCI200524408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Li Y, Huang TT, Carlson EJ, Melov S, Ursell PC, Olson JL, et al. Dilated cardiomyopathy and neonatal lethality in mutant mice lacking manganese superoxide dismutase. Nat Genet. 1995;11:376–81. doi: 10.1038/ng1295-376. [DOI] [PubMed] [Google Scholar]
  • 40.Gladyshev VN, Liu A, Novoselov SV, Krysan K, Sun QA, Kryukov VM, et al. Identification and characterization of a new mammalian glutaredoxin (thioltransferase), Grx2. J Biol Chem. 2001;276:30374–80. doi: 10.1074/jbc.M100020200. [DOI] [PubMed] [Google Scholar]
  • 41.Schriner SE, Linford NJ, Martin GM, Treuting P, Ogburn CE, Emond M, et al. Extension of murine life span by overexpression of catalase targeted to mitochondria. Science. 2005;308:1909–11. doi: 10.1126/science.1106653. [DOI] [PubMed] [Google Scholar]
  • 42.Matsui M, Oshima M, Oshima H, Takaku K, Maruyama T, Yodoi J, et al. Early embryonic lethality caused by targeted disruption of the mouse thioredoxin gene. Dev Biol. 1996;178:179–85. doi: 10.1006/dbio.1996.0208. [DOI] [PubMed] [Google Scholar]
  • 43.Jakupoglu C, Przemeck GK, Schneider M, Moreno SG, Mayr N, Hatzopoulos AK, et al. Cytoplasmic thioredoxin reductase is essential for embryogenesis but dispensable for cardiac development. Mol Cell Biol. 2005;25:1980–8. doi: 10.1128/MCB.25.5.1980-1988.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Tanaka T, Hosoi F, Yamaguchi-Iwai Y, Nakamura H, Masutani H, Ueda S, et al. Thioredoxin-2 (TRX-2) is an essential gene regulating mitochondria-dependent apoptosis. Embo J. 2002;21:1695–703. doi: 10.1093/emboj/21.7.1695. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Nonn L, Williams RR, Erickson RP, Powis G. The absence of mitochondrial thioredoxin 2 causes massive apoptosis, exencephaly, and early embryonic lethality in homozygous mice. Mol Cell Biol. 2003;23:916–22. doi: 10.1128/MCB.23.3.916-922.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Conrad M, Jakupoglu C, Moreno SG, Lippl S, Banjac A, Schneider M, et al. Essential role for mitochondrial thioredoxin reductase in hematopoiesis, heart development, and heart function. Mol Cell Biol. 2004;24:9414–23. doi: 10.1128/MCB.24.21.9414-9423.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Mann DL. Left ventricular size and shape: determinants of mechanical signal transduction pathways. Heart Fail Rev. 2005;10:95–100. doi: 10.1007/s10741-005-4636-y. [DOI] [PubMed] [Google Scholar]
  • 48.van Bilsen M, Smeets PJ, Gilde AJ, van der Vusse GJ. Metabolic remodelling of the failing heart: the cardiac burn-out syndrome? Cardiovasc Res. 2004;61:218–26. doi: 10.1016/j.cardiores.2003.11.014. [DOI] [PubMed] [Google Scholar]
  • 49.Barger PM, Kelly DP. PPAR signaling in the control of cardiac energy metabolism. Trends Cardiovasc Med. 2000;10:238–45. doi: 10.1016/s1050-1738(00)00077-3. [DOI] [PubMed] [Google Scholar]
  • 50.Ikeuchi M, Tsutsui H, Shiomi T, Matsusaka H, Matsushima S, Wen J, et al. Inhibition of TGF-beta signaling exacerbates early cardiac dysfunction but prevents late remodeling after infarction. Cardiovasc Res. 2004;64:526–35. doi: 10.1016/j.cardiores.2004.07.017. [DOI] [PubMed] [Google Scholar]
  • 51.Opie LH, Commerford PJ, Gersh BJ, Pfeffer MA. Controversies in ventricular remodelling. Lancet. 2006;367:356–67. doi: 10.1016/S0140-6736(06)68074-4. [DOI] [PubMed] [Google Scholar]
  • 52.Mann DL, Bristow MR. Mechanisms and models in heart failure: the biomechanical model and beyond. Circulation. 2005;111:2837–49. doi: 10.1161/CIRCULATIONAHA.104.500546. [DOI] [PubMed] [Google Scholar]
  • 53.Byrne JA, Grieve DJ, Cave AC, Shah AM. Oxidative stress and heart failure. Arch Mal Coeur Vaiss. 2003;96:214–21. [PubMed] [Google Scholar]
  • 54.Grieve DJ, Byrne JA, Cave AC, Shah AM. Role of oxidative stress in cardiac remodelling after myocardial infarction. Heart Lung Circ. 2004;13:132–8. doi: 10.1016/j.hlc.2004.02.008. [DOI] [PubMed] [Google Scholar]
  • 55.Murdoch CE, Zhang M, Cave AC, Shah AM. NADPH oxidase-dependent redox signalling in cardiac hypertrophy, remodelling and failure. Cardiovasc Res. 2006 doi: 10.1016/j.cardiores.2006.03.016. [DOI] [PubMed] [Google Scholar]
  • 56.Amin JK, Xiao L, Pimental DR, Pagano PJ, Singh K, Sawyer DB, et al. Reactive oxygen species mediate alpha-adrenergic receptor-stimulated hypertrophy in adult rat ventricular myocytes. J Mol Cell Cardiol. 2001;33:131–9. doi: 10.1006/jmcc.2000.1285. [DOI] [PubMed] [Google Scholar]
  • 57.Mollnau H, Wendt M, Szocs K, Lassegue B, Schulz E, Oelze M, et al. Effects of angiotensin II infusion on the expression and function of NAD(P)H oxidase and components of nitric oxide/cGMP signaling. Circ Res. 2002;90:E58–65. doi: 10.1161/01.res.0000012569.55432.02. [DOI] [PubMed] [Google Scholar]
  • 58.Pimentel DR, Amin JK, Xiao L, Miller T, Viereck J, Oliver-Krasinski J, et al. Reactive oxygen species mediate amplitude-dependent hypertrophic and apoptotic responses to mechanical stretch in cardiac myocytes. Circ Res. 2001;89:453–60. doi: 10.1161/hh1701.096615. [DOI] [PubMed] [Google Scholar]
  • 59.Ide T, Tsutsui H, Kinugawa S, Utsumi H, Kang D, Hattori N, et al. Mitochondrial electron transport complex I is a potential source of oxygen free radicals in the failing myocardium. Circ Res. 1999;85:357–63. doi: 10.1161/01.res.85.4.357. [DOI] [PubMed] [Google Scholar]
  • 60.Sayen MR, Gustafsson AB, Sussman MA, Molkentin JD, Gottlieb RA. Calcineurin transgenic mice have mitochondrial dysfunction and elevated superoxide production. Am J Physiol Cell Physiol. 2003;284:C562–70. doi: 10.1152/ajpcell.00336.2002. [DOI] [PubMed] [Google Scholar]
  • 61.Li JM, Gall NP, Grieve DJ, Chen M, Shah AM. Activation of NADPH oxidase during progression of cardiac hypertrophy to failure. Hypertension. 2002;40:477–84. doi: 10.1161/01.hyp.0000032031.30374.32. [DOI] [PubMed] [Google Scholar]
  • 62.Sorescu D, Griendling KK. Reactive oxygen species, mitochondria, and NAD(P)H oxidases in the development and progression of heart failure. Congest Heart Fail. 2002;8:132–40. doi: 10.1111/j.1527-5299.2002.00717.x. [DOI] [PubMed] [Google Scholar]
  • 63.Xiao L, Pimentel DR, Wang J, Singh K, Colucci WS, Sawyer DB. Role of reactive oxygen species and NAD(P)H oxidase in alpha(1)-adrenoceptor signaling in adult rat cardiac myocytes. Am J Physiol Cell Physiol. 2002;282:C926–34. doi: 10.1152/ajpcell.00254.2001. [DOI] [PubMed] [Google Scholar]
  • 64.Saavedra WF, Paolocci N, St John ME, Skaf MW, Stewart GC, Xie JS, et al. Imbalance between xanthine oxidase and nitric oxide synthase signaling pathways underlies mechanoenergetic uncoupling in the failing heart. Circ Res. 2002;90:297–304. doi: 10.1161/hh0302.104531. [DOI] [PubMed] [Google Scholar]
  • 65.Landmesser U, Dikalov S, Price SR, McCann L, Fukai T, Holland SM, et al. Oxidation of tetrahydrobiopterin leads to uncoupling of endothelial cell nitric oxide synthase in hypertension. J Clin Invest. 2003;111:1201–9. doi: 10.1172/JCI14172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Takimoto E, Champion HC, Li M, Ren S, Rodriguez ER, Tavazzi B, et al. Oxidant stress from nitric oxide synthase-3 uncoupling stimulates cardiac pathologic remodeling from chronic pressure load. J Clin Invest. 2005;115:1221–31. doi: 10.1172/JCI21968. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Fosslien E. Review: Mitochondrial medicine--cardiomyopathy caused by defective oxidative phosphorylation. Ann Clin Lab Sci. 2003;33:371–95. [PubMed] [Google Scholar]
  • 68.Yodoi J, Tursz T. ADF, a growth-promoting factor derived from adult T cell leukemia and homologous to thioredoxin: involvement in lymphocyte immortalization by HTLV-I and EBV. Adv Cancer Res. 1991;57:381–411. doi: 10.1016/s0065-230x(08)61004-0. [DOI] [PubMed] [Google Scholar]
  • 69.Kuster GM, Pimentel DR, Adachi T, Ido Y, Brenner DA, Cohen RA, et al. Alpha-adrenergic receptor-stimulated hypertrophy in adult rat ventricular myocytes is mediated via thioredoxin-1-sensitive oxidative modification of thiols on Ras. Circulation. 2005;111:1192–8. doi: 10.1161/01.CIR.0000157148.59308.F5. [DOI] [PubMed] [Google Scholar]
  • 70.Bolli R. The late phase of preconditioning. Circ Res. 2000;87:972–83. doi: 10.1161/01.res.87.11.972. [DOI] [PubMed] [Google Scholar]
  • 71.Murry CE, Jennings RB, Reimer KA. Preconditioning with ischemia: a delay of lethal cell injury in ischemic myocardium. Circulation. 1986;74:1124–36. doi: 10.1161/01.cir.74.5.1124. [DOI] [PubMed] [Google Scholar]
  • 72.Das DK. Thioredoxin regulation of ischemic preconditioning. Antioxid Redox Signal. 2004;6:405–12. doi: 10.1089/152308604322899477. [DOI] [PubMed] [Google Scholar]
  • 73.Gray MO, Zhou HZ, Schafhalter-Zoppoth I, Zhu P, Mochly-Rosen D, Messing RO. Preservation of base-line hemodynamic function and loss of inducible cardioprotection in adult mice lacking protein kinase C epsilon. J Biol Chem. 2004;279:3596–604. doi: 10.1074/jbc.M311459200. [DOI] [PubMed] [Google Scholar]
  • 74.Bell RM, Cave AC, Johar S, Hearse DJ, Shah AM, Shattock MJ. Pivotal role of NOX-2-containing NADPH oxidase in early ischemic preconditioning. Faseb J. 2005;19:2037–9. doi: 10.1096/fj.04-2774fje. [DOI] [PubMed] [Google Scholar]
  • 75.Mayr M, Metzler B, Chung YL, McGregor E, Mayr U, Troy H, et al. Ischemic preconditioning exaggerates cardiac damage in PKC-delta null mice. Am J Physiol Heart Circ Physiol. 2004;287:H946–56. doi: 10.1152/ajpheart.00878.2003. [DOI] [PubMed] [Google Scholar]
  • 76.Ping P, Song C, Zhang J, Guo Y, Cao X, Li RC, et al. Formation of protein kinase C(epsilon)-Lck signaling modules confers cardioprotection. J Clin Invest. 2002;109:499–507. doi: 10.1172/JCI13200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Das DK, Maulik N. Preconditioning potentiates redox signaling and converts death signal into survival signal. Arch Biochem Biophys. 2003;420:305–11. doi: 10.1016/j.abb.2003.09.023. [DOI] [PubMed] [Google Scholar]
  • 78.Malik G, Gorvanov N, Das S, Narasimhan G, Otani H, Maulik N, et al. Ischemic proconditioning triggers nuclear translocation of thioredoxin and its interaction with ref-1 poteintiating a survival signal through PI-3-kinase-akt pathway. Antioxid Redox Signal. doi: 10.1089/ars.2006.8.2101. in press. [DOI] [PubMed] [Google Scholar]
  • 79.Soejima H, Suefuji H, Miyamoto S, Kajiwaram I, Kojima S, Hokamaki J, et al. Increased plasma thioredoxin in patients with acute myocardial infarction. Clin Cardiol. 2003;26:583–7. doi: 10.1002/clc.4960261208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Aota M, Matsuda K, Isowa N, Wada H, Yodoi J, Ban T. Protection against reperfusion-induced arrhythmias by human thioredoxin. J Cardiovasc Pharmacol. 1996;27:727–32. doi: 10.1097/00005344-199605000-00016. [DOI] [PubMed] [Google Scholar]
  • 81.Francis GS. Pathophysiology of chronic heart failure. Am J Med. 2001;110(Suppl 7A):37S–46S. doi: 10.1016/s0002-9343(98)00385-4. [DOI] [PubMed] [Google Scholar]
  • 82.Scarabelli TM, Gottlieb RA. Functional and clinical repercussions of myocyte apoptosis in the multifaceted damage by ischemia/reperfusion injury: old and new concepts after 10 years of contributions. Cell Death Differ. 2004;11 (Suppl 2):S144–52. doi: 10.1038/sj.cdd.4401544. [DOI] [PubMed] [Google Scholar]
  • 83.Kourie JI. Interaction of reactive oxygen species with ion transport mechanisms. Am J Physiol. 1998;275:C1–24. doi: 10.1152/ajpcell.1998.275.1.C1. [DOI] [PubMed] [Google Scholar]
  • 84.Guerra L, Cerbai E, Gessi S, Borea PA, Mugelli A. The effect of oxygen free radicals on calcium current and dihydropyridine binding sites in guinea-pig ventricular myocytes. Br J Pharmacol. 1996;118:1278–84. doi: 10.1111/j.1476-5381.1996.tb15534.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Kaplan P, Babusikova E, Lehotsky J, Dobrota D. Free radical-induced protein modification and inhibition of Ca2+-ATPase of cardiac sarcoplasmic reticulum. Mol Cell Biochem. 2003;248:41–7. doi: 10.1023/a:1024145212616. [DOI] [PubMed] [Google Scholar]
  • 86.Rohrbach S, Niemann B, Abushouk AM, Holtz J. Caloric restriction and mitochondrial function in the ageing myocardium. Exp Gerontol. 2006 doi: 10.1016/j.exger.2006.02.001. [DOI] [PubMed] [Google Scholar]
  • 87.Kishimoto C, Shioji K, Nakamura H, Nakayama Y, Yodoi J, Sasayama S. Serum thioredoxin (TRX) levels in patients with heart failure. Jpn Circ J. 2001;65:491–4. doi: 10.1253/jcj.65.491. [DOI] [PubMed] [Google Scholar]
  • 88.Jekell A, Hossain A, Alehagen U, Dahlstrom U, Rosen A. Elevated circulating levels of thioredoxin and stress in chronic heart failure. Eur J Heart Fail. 2004;6:883–90. doi: 10.1016/j.ejheart.2004.03.003. [DOI] [PubMed] [Google Scholar]
  • 89.Liu W, Nakamura H, Shioji K, Tanito M, Oka S, Ahsan MK, et al. Thioredoxin-1 ameliorates myosin-induced autoimmune myocarditis by suppressing chemokine expressions and leukocyte chemotaxis in mice. Circulation. 2004;110:1276–83. doi: 10.1161/01.CIR.0000141803.41217.B6. [DOI] [PubMed] [Google Scholar]
  • 90.Yamamoto S, Yang G, Zablocki D, Liu J, Hong C, Kim SJ, et al. Activation of Mst1 causes dilated cardiomyopathy by stimulating apoptosis without compensatory ventricular myocyte hypertrophy. J Clin Invest. 2003;111:1463–74. doi: 10.1172/JCI17459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Saitoh M, Nishitoh H, Fujii M, Takeda K, Tobiume K, Sawada Y, et al. Mammalian thioredoxin is a direct inhibitor of apoptosis signal-regulating kinase (ASK) 1. Embo J. 1998;17:2596–606. doi: 10.1093/emboj/17.9.2596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Liu Y, Min W. Thioredoxin promotes ASK1 ubiquitination and degradation to inhibit ASK1-mediated apoptosis in a redox activity-independent manner. Circ Res. 2002;90:1259–66. doi: 10.1161/01.res.0000022160.64355.62. [DOI] [PubMed] [Google Scholar]
  • 93.Zhang R, Al-Lamki R, Bai L, Streb JW, Miano JM, Bradley J, et al. Thioredoxin-2 inhibits mitochondria-located ASK1-mediated apoptosis in a JNK-independent manner. Circ Res. 2004;94:1483–91. doi: 10.1161/01.RES.0000130525.37646.a7. [DOI] [PubMed] [Google Scholar]
  • 94.Nishida K, Otsu K. The role of apoptosis signal-regulating kinase 1 in cardiomyocyte apoptosis. Antioxid Redox Signal. doi: 10.1089/ars.2006.8.1729. in press. [DOI] [PubMed] [Google Scholar]
  • 95.Meuillet EJ, Mahadevan D, Berggren M, Coon A, Powis G. Thioredoxin-1 binds to the C2 domain of PTEN inhibiting PTEN's lipid phosphatase activity and membrane binding: a mechanism for the functional loss of PTEN's tumor suppressor activity. Arch Biochem Biophys. 2004;429:123–33. doi: 10.1016/j.abb.2004.04.020. [DOI] [PubMed] [Google Scholar]
  • 96.Kaimul Ahsan M, Nakamura H, Tanito M, Yamada K, Utsumi H, Yodoi J. Thioredoxin-1 suppresses lung injury and apoptosis induced by diesel exhaust particles (DEP) by scavenging reactive oxygen species and by inhibiting DEP-induced downregulation of Akt. Free Radic Biol Med. 2005;39:1549–59. doi: 10.1016/j.freeradbiomed.2005.07.016. [DOI] [PubMed] [Google Scholar]
  • 97.Datta SR, Dudek H, Tao X, Masters S, Fu H, Gotoh Y, et al. Akt phosphorylation of BAD couples survival signals to the cell-intrinsic death machinery. Cell. 1997;91:231–41. doi: 10.1016/s0092-8674(00)80405-5. [DOI] [PubMed] [Google Scholar]
  • 98.Cardone MH, Roy N, Stennicke HR, Salvesen GS, Franke TF, Stanbridge E, et al. Regulation of cell death protease caspase-9 by phosphorylation. Science. 1998;282:1318–21. doi: 10.1126/science.282.5392.1318. [DOI] [PubMed] [Google Scholar]
  • 99.Dawn B, Xuan YT, Marian M, Flaherty MP, Murphree SS, Smith TL, et al. Cardiac-specific abrogation of NF- kappa B activation in mice by transdominant expression of a mutant I kappa B alpha. J Mol Cell Cardiol. 2001;33:161–73. doi: 10.1006/jmcc.2000.1291. [DOI] [PubMed] [Google Scholar]
  • 100.Spallarossa P, Altieri P, Garibaldi S, Ghigliotti G, Barisione C, Manca V, et al. Matrix metalloproteinase-2 and -9 are induced differently by doxorubicin in H9c2 cells: The role of MAP kinases and NAD(P)H oxidase. Cardiovasc Res. 2006;69:736–45. doi: 10.1016/j.cardiores.2005.08.009. [DOI] [PubMed] [Google Scholar]
  • 101.Siwik DA, Pagano PJ, Colucci WS. Oxidative stress regulates collagen synthesis and matrix metalloproteinase activity in cardiac fibroblasts. Am J Physiol Cell Physiol. 2001;280:C53–60. doi: 10.1152/ajpcell.2001.280.1.C53. [DOI] [PubMed] [Google Scholar]
  • 102.Siwik DA, Colucci WS. Regulation of matrix metalloproteinases by cytokines and reactive oxygen/nitrogen species in the myocardium. Heart Fail Rev. 2004;9:43–51. doi: 10.1023/B:HREV.0000011393.40674.13. [DOI] [PubMed] [Google Scholar]
  • 103.Cucoranu I, Clempus R, Dikalova A, Phelan PJ, Ariyan S, Dikalov S, et al. NAD(P)H oxidase 4 mediates transforming growth factor-beta1-induced differentiation of cardiac fibroblasts into myofibroblasts. Circ Res. 2005;97:900–7. doi: 10.1161/01.RES.0000187457.24338.3D. [DOI] [PubMed] [Google Scholar]
  • 104.Hoshino T, Nakamura H, Okamoto M, Kato S, Araya S, Nomiyama K, et al. Redox-active protein thioredoxin prevents proinflammatory cytokine- or bleomycin-induced lung injury. Am J Respir Crit Care Med. 2003;168:1075–83. doi: 10.1164/rccm.200209-982OC. [DOI] [PubMed] [Google Scholar]
  • 105.Yoshida T, Oka S, Masutani H, Nakamura H, Yodoi J. The role of thioredoxin in the aging process: involvement of oxidative stress. Antioxid Redox Signal. 2003;5:563–70. doi: 10.1089/152308603770310211. [DOI] [PubMed] [Google Scholar]
  • 106.Mitsui A, Hamuro J, Nakamura H, Kondo N, Hirabayashi Y, Ishizaki-Koizumi S, et al. Overexpression of human thioredoxin in transgenic mice controls oxidative stress and life span. Antioxid Redox Signal. 2002;4:693–6. doi: 10.1089/15230860260220201. [DOI] [PubMed] [Google Scholar]
  • 107.Yoshida T, Nakamura H, Masutani H, Yodoi J. The involvement of thioredoxin and thioredoxin binding protein-2 on cellular proliferation and aging process. Ann N Y Acad Sci. 2005;1055:1–12. doi: 10.1196/annals.1323.002. [DOI] [PubMed] [Google Scholar]
  • 108.Rohrbach S, Gruenler S, Teschner M, Holtz J. The thioredoxin system in aging muscle: key role of mitochondrial thioredoxin reductase in the protective effects of caloric restriction? Am J Physiol Regul Integr Comp Physiol. 2006 doi: 10.1152/ajpregu.00890.2005. [DOI] [PubMed] [Google Scholar]
  • 109.Arner ES, Holmgren A. Physiological functions of thioredoxin and thioredoxin reductase. Eur J Biochem. 2000;267:6102–9. doi: 10.1046/j.1432-1327.2000.01701.x. [DOI] [PubMed] [Google Scholar]
  • 110.Petropoulos I, Friguet B. Protein maintenance in aging and replicative senescence: a role for the peptide methionine sulfoxide reductases. Biochim Biophys Acta. 2005;1703:261–6. doi: 10.1016/j.bbapap.2004.08.018. [DOI] [PubMed] [Google Scholar]
  • 111.Kuschel L, Hansel A, Schonherr R, Weissbach H, Brot N, Hoshi T, et al. Molecular cloning and functional expression of a human peptide methionine sulfoxide reductase (hMsrA) FEBS Lett. 1999;456:17–21. doi: 10.1016/s0014-5793(99)00917-5. [DOI] [PubMed] [Google Scholar]
  • 112.Moskovitz J, Bar-Noy S, Williams WM, Requena J, Berlett BS, Stadtman ER. Methionine sulfoxide reductase (MsrA) is a regulator of antioxidant defense and lifespan in mammals. Proc Natl Acad Sci U S A. 2001;98:12920–5. doi: 10.1073/pnas.231472998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Stadtman ER, Van Remmen H, Richardson A, Wehr NB, Levine RL. Methionine oxidation and aging. Biochim Biophys Acta. 2005;1703:135–40. doi: 10.1016/j.bbapap.2004.08.010. [DOI] [PubMed] [Google Scholar]
  • 114.Jafri MS, Dudycha SJ, O'Rourke B. Cardiac energy metabolism: models of cellular respiration. Annu Rev Biomed Eng. 2001;3:57–81. doi: 10.1146/annurev.bioeng.3.1.57. [DOI] [PubMed] [Google Scholar]
  • 115.Ritchie RH, Delbridge LM. Cardiac hypertrophy, substrate utilization and metabolic remodelling: cause or effect? Clin Exp Pharmacol Physiol. 2006;33:159–66. doi: 10.1111/j.1440-1681.2006.04342.x. [DOI] [PubMed] [Google Scholar]
  • 116.Finck BN, Kelly DP. PGC-1 coactivators: inducible regulators of energy metabolism in health and disease. J Clin Invest. 2006;116:615–22. doi: 10.1172/JCI27794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Huss JM, Kelly DP. Nuclear receptor signaling and cardiac energetics. Circ Res. 2004;95:568–78. doi: 10.1161/01.RES.0000141774.29937.e3. [DOI] [PubMed] [Google Scholar]
  • 118.Sano M, Wang SC, Shirai M, Scaglia F, Xie M, Sakai S, et al. Activation of cardiac Cdk9 represses PGC-1 and confers a predisposition to heart failure. Embo J. 2004;23:3559–69. doi: 10.1038/sj.emboj.7600351. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Huss JM, Kelly DP. Mitochondrial energy metabolism in heart failure: a question of balance. J Clin Invest. 2005;115:547–55. doi: 10.1172/JCI200524405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Cheng L, Ding G, Qin Q, Huang Y, Lewis W, He N, et al. Cardiomyocyte-restricted peroxisome proliferator-activated receptor-delta deletion perturbs myocardial fatty acid oxidation and leads to cardiomyopathy. Nat Med. 2004;10:1245–50. doi: 10.1038/nm1116. [DOI] [PubMed] [Google Scholar]
  • 121.Arany Z, Novikov M, Chin S, Ma Y, Rosenzweig A, Spiegelman BM. Transverse aortic constriction leads to accelerated heart failure in mice lacking PPAR-{gamma} coactivator 1{alpha} Proc Natl Acad Sci U S A. 2006;103:10086–91. doi: 10.1073/pnas.0603615103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Luptak I, Balschi JA, Xing Y, Leone TC, Kelly DP, Tian R. Decreased contractile and metabolic reserve in peroxisome proliferator-activated receptor-alpha-null hearts can be rescued by increasing glucose transport and utilization. Circulation. 2005;112:2339–46. doi: 10.1161/CIRCULATIONAHA.105.534594. [DOI] [PubMed] [Google Scholar]
  • 123.Arany Z, He H, Lin J, Hoyer K, Handschin C, Toka O, et al. Transcriptional coactivator PGC-1 alpha controls the energy state and contractile function of cardiac muscle. Cell Metab. 2005;1:259–71. doi: 10.1016/j.cmet.2005.03.002. [DOI] [PubMed] [Google Scholar]
  • 124.Barger PM, Kelly DP. Fatty acid utilization in the hypertrophied and failing heart: molecular regulatory mechanisms. Am J Med Sci. 1999;318:36–42. doi: 10.1097/00000441-199907000-00006. [DOI] [PubMed] [Google Scholar]
  • 125.Ago T, Yeh I, Yamamoto M, Shinke-Braun M, Brown JA, Tian B, et al. Thioredoxin1 upregulates mitochondrial proteins related to oxidative stress and TCA cycle in the heart. Antioxid Redox Signal. doi: 10.1089/ars.2006.8.1635. in press. [DOI] [PubMed] [Google Scholar]
  • 126.Hodgson DM, Zingman LV, Kane GC, Perez-Terzic C, Bienengraeber M, Ozcan C, et al. Cellular remodeling in heart failure disrupts K(ATP) channel-dependent stress tolerance. Embo J. 2003;22:1732–42. doi: 10.1093/emboj/cdg192. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Kane GC, Liu XK, Yamada S, Olson TM, Terzic A. Cardiac KATP channels in health and disease. J Mol Cell Cardiol. 2005;38:937–43. doi: 10.1016/j.yjmcc.2005.02.026. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Li X, Li S, Xu Z, Lou MF, Anding P, Liu D, et al. Redox control of K+ channel remodeling in rat ventricle. J Mol Cell Cardiol. 2006;40:339–49. doi: 10.1016/j.yjmcc.2005.09.019. [DOI] [PubMed] [Google Scholar]
  • 129.Ichijo H, Nishida E, Irie K, ten Dijke P, Saitoh M, Moriguchi T, et al. Induction of apoptosis by ASK1, a mammalian MAPKKK that activates SAPK/JNK and p38 signaling pathways. Science. 1997;275:90–4. doi: 10.1126/science.275.5296.90. [DOI] [PubMed] [Google Scholar]
  • 130.Hirotani S, Otsu K, Nishida K, Higuchi Y, Morita T, Nakayama H, et al. Involvement of nuclear factor-kappaB and apoptosis signal-regulating kinase 1 in G-protein-coupled receptor agonist-induced cardiomyocyte hypertrophy. Circulation. 2002;105:509–15. doi: 10.1161/hc0402.102863. [DOI] [PubMed] [Google Scholar]
  • 131.Izumiya Y, Kim S, Izumi Y, Yoshida K, Yoshiyama M, Matsuzawa A, et al. Apoptosis signal-regulating kinase 1 plays a pivotal role in angiotensin II-induced cardiac hypertrophy and remodeling. Circ Res. 2003;93:874–83. doi: 10.1161/01.RES.0000100665.67510.F5. [DOI] [PubMed] [Google Scholar]
  • 132.He X, Liu Y, Sharma V, Dirksen RT, Waugh R, Sheu SS, et al. ASK1 associates with troponin T and induces troponin T phosphorylation and contractile dysfunction in cardiomyocytes. Am J Pathol. 2003;163:243–51. doi: 10.1016/S0002-9440(10)63647-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Chen KS, DeLuca HF. Isolation and characterization of a novel cDNA from HL-60 cells treated with 1,25-dihydroxyvitamin D-3. Biochim Biophys Acta. 1994;1219:26–32. doi: 10.1016/0167-4781(94)90242-9. [DOI] [PubMed] [Google Scholar]
  • 134.Oka S, Liu W, Masutani H, Hirata H, Shinkai Y, Yamada S, et al. Impaired fatty acid utilization in thioredoxin binding protein-2 (TBP-2)-deficient mice: a unique animal model of Reye syndrome. Faseb J. 2006;20:121–3. doi: 10.1096/fj.05-4439fje. [DOI] [PubMed] [Google Scholar]
  • 135.Wang Y, De Keulenaer GW, Lee RT. Vitamin D(3)-up-regulated protein-1 is a stress-responsive gene that regulates cardiomyocyte viability through interaction with thioredoxin. J Biol Chem. 2002;277:26496–500. doi: 10.1074/jbc.M202133200. [DOI] [PubMed] [Google Scholar]
  • 136.Yoshioka J, Schulze PC, Cupesi M, Sylvan JD, MacGillivray C, Gannon J, et al. Thioredoxin-interacting protein controls cardiac hypertrophy through regulation of thioredoxin activity. Circulation. 2004;109:2581–6. doi: 10.1161/01.CIR.0000129771.32215.44. [DOI] [PubMed] [Google Scholar]
  • 137.Han SH, Jeon JH, Ju HR, Jung U, Kim KY, Yoo HS, et al. VDUP1 upregulated by TGF-beta1 and 1,25-dihydorxyvitamin D3 inhibits tumor cell growth by blocking cell-cycle progression. Oncogene. 2003;22:4035–46. doi: 10.1038/sj.onc.1206610. [DOI] [PubMed] [Google Scholar]
  • 138.Kim KY, Shin SM, Kim JK, Paik SG, Yang Y, Choi I. Heat shock factor regulates VDUP1 gene expression. Biochem Biophys Res Commun. 2004;315:369–75. doi: 10.1016/j.bbrc.2004.01.047. [DOI] [PubMed] [Google Scholar]
  • 139.Bodnar JS, Chatterjee A, Castellani LW, Ross DA, Ohmen J, Cavalcoli J, et al. Positional cloning of the combined hyperlipidemia gene Hyplip1. Nat Genet. 2002;30:110–6. doi: 10.1038/ng811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Seo JH, Ahn Y, Lee SR, Yeol Yeo C, Chung Hur K. The major target of the endogenously generated reactive oxygen species in response to insulin stimulation is phosphatase and tensin homolog and not phosphoinositide-3 kinase (PI-3 kinase) in the PI-3 kinase/Akt pathway. Mol Biol Cell. 2005;16:348–57. doi: 10.1091/mbc.E04-05-0369. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Cai Z, Semenza GL. PTEN activity is modulated during ischemia and reperfusion: involvement in the induction and decay of preconditioning. Circ Res. 2005;97:1351–9. doi: 10.1161/01.RES.0000195656.52760.30. [DOI] [PubMed] [Google Scholar]
  • 142.Wakatsuki T, Schlessinger J, Elson EL. The biochemical response of the heart to hypertension and exercise. Trends Biochem Sci. 2004;29:609–17. doi: 10.1016/j.tibs.2004.09.002. [DOI] [PubMed] [Google Scholar]
  • 143.DeBosch B, Treskov I, Lupu TS, Weinheimer C, Kovacs A, Courtois M, et al. Akt1 is required for physiological cardiac growth. Circulation. 2006;113:2097–104. doi: 10.1161/CIRCULATIONAHA.105.595231. [DOI] [PubMed] [Google Scholar]
  • 144.Matthews JR, Wakasugi N, Virelizier JL, Yodoi J, Hay RT. Thioredoxin regulates the DNA binding activity of NF-kappa B by reduction of a disulphide bond involving cysteine 62. Nucleic Acids Res. 1992;20:3821–30. doi: 10.1093/nar/20.15.3821. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Hirota K, Murata M, Sachi Y, Nakamura H, Takeuchi J, Mori K, et al. Distinct roles of thioredoxin in the cytoplasm and in the nucleus. A two-step mechanism of redox regulation of transcription factor NF-kappaB. J Biol Chem. 1999;274:27891–7. doi: 10.1074/jbc.274.39.27891. [DOI] [PubMed] [Google Scholar]
  • 146.Qin J, Clore GM, Kennedy WP, Kuszewski J, Gronenborn AM. The solution structure of human thioredoxin complexed with its target from Ref-1 reveals peptide chain reversal. Structure. 1996;4:613–20. doi: 10.1016/s0969-2126(96)00065-2. [DOI] [PubMed] [Google Scholar]
  • 147.Hirota K, Matsui M, Iwata S, Nishiyama A, Mori K, Yodoi J. AP-1 transcriptional activity is regulated by a direct association between thioredoxin and Ref-1. Proc Natl Acad Sci U S A. 1997;94:3633–8. doi: 10.1073/pnas.94.8.3633. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Purcell NH, Tang G, Yu C, Mercurio F, DiDonato JA, Lin A. Activation of NF-kappa B is required for hypertrophic growth of primary rat neonatal ventricular cardiomyocytes. Proc Natl Acad Sci U S A. 2001;98:6668–73. doi: 10.1073/pnas.111155798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Kovacic-Milivojevic B, Wong VS, Gardner DG. Selective regulation of the atrial natriuretic peptide gene by individual components of the activator protein-1 complex. Endocrinology. 1996;137:1108–17. doi: 10.1210/endo.137.3.8603581. [DOI] [PubMed] [Google Scholar]
  • 150.Hansen JM, Watson WH, Jones DP. Compartmentation of Nrf-2 redox control: regulation of cytoplasmic activation by glutathione and DNA binding by thioredoxin-1. Toxicol Sci. 2004;82:308–17. doi: 10.1093/toxsci/kfh231. [DOI] [PubMed] [Google Scholar]
  • 151.Hardt SE, Sadoshima J. Negative regulators of cardiac hypertrophy. Cardiovasc Res. 2004;63:500–9. doi: 10.1016/j.cardiores.2004.03.015. [DOI] [PubMed] [Google Scholar]
  • 152.Tomita H, Nazmy M, Kajimoto K, Yehia G, Molina CA, Sadoshima J. Inducible cAMP early repressor (ICER) is a negative-feedback regulator of cardiac hypertrophy and an important mediator of cardiac myocyte apoptosis in response to beta-adrenergic receptor stimulation. Circ Res. 2003;93:12–22. doi: 10.1161/01.RES.0000079794.57578.F1. [DOI] [PubMed] [Google Scholar]
  • 153.Kaga S, Zhan L, Matsumoto M, Maulik N. Resveratrol enhances neovascularization in the infarcted rat myocardium through the induction of thioredoxin-1, heme oxygenase-1 and vascular endothelial growth factor. J Mol Cell Cardiol. 2005;39:813–22. doi: 10.1016/j.yjmcc.2005.08.003. [DOI] [PubMed] [Google Scholar]

RESOURCES