Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Endurance exercise stimulates peroxisome proliferator-activated receptor gamma coactivator-1alpha (PGC-1alpha) expression in skeletal muscle, and forced expression of PGC-1alpha changes muscle metabolism and exercise capacity in mice. However, it is unclear if PGC-1alpha is indispensible for endurance exercise-induced metabolic and contractile adaptations in skeletal muscle. In this study, we showed that endurance exercise-induced expression of mitochondrial enzymes (cytochrome oxidase IV and cytochrome c) and increases of platelet endothelial cell adhesion molecule-1 (PECAM-1, CD31)-positive endothelial cells in skeletal muscle, but not IIb-to-IIa fiber-type transformation, were significantly attenuated in muscle-specific Pgc-1alpha knockout mice. Interestingly, voluntary running effectively restored the compromised mitochondrial integrity and superoxide dismutase 2 (SOD2) protein expression in skeletal muscle in Pgc-1alpha knockout mice. Thus, PGC-1alpha plays a functional role in endurance exercise-induced mitochondrial biogenesis and angiogenesis, but not IIb-to-IIa fiber-type transformation in mouse skeletal muscle, and the improvement of mitochondrial morphology and antioxidant defense in response to endurance exercise may occur independently of PGC-1alpha function. We conclude that PGC-1alpha is required for complete skeletal muscle adaptations induced by endurance exercise in mice.

Free full text 


Logo of ajpcellLink to Publisher's site
Am J Physiol Cell Physiol. 2010 Mar; 298(3): C572–C579.
Published online 2009 Dec 23. https://doi.org/10.1152/ajpcell.00481.2009
PMCID: PMC3353735
PMID: 20032509

PGC-1α plays a functional role in exercise-induced mitochondrial biogenesis and angiogenesis but not fiber-type transformation in mouse skeletal muscle

Associated Data

Supplementary Materials

Abstract

Endurance exercise stimulates peroxisome proliferator-activated receptor γ coactivator-1α (PGC-1α) expression in skeletal muscle, and forced expression of PGC-1α changes muscle metabolism and exercise capacity in mice. However, it is unclear if PGC-1α is indispensible for endurance exercise-induced metabolic and contractile adaptations in skeletal muscle. In this study, we showed that endurance exercise-induced expression of mitochondrial enzymes (cytochrome oxidase IV and cytochrome c) and increases of platelet endothelial cell adhesion molecule-1 (PECAM-1, CD31)-positive endothelial cells in skeletal muscle, but not IIb-to-IIa fiber-type transformation, were significantly attenuated in muscle-specific Pgc-1α knockout mice. Interestingly, voluntary running effectively restored the compromised mitochondrial integrity and superoxide dismutase 2 (SOD2) protein expression in skeletal muscle in Pgc-1α knockout mice. Thus, PGC-1α plays a functional role in endurance exercise-induced mitochondrial biogenesis and angiogenesis, but not IIb-to-IIa fiber-type transformation in mouse skeletal muscle, and the improvement of mitochondrial morphology and antioxidant defense in response to endurance exercise may occur independently of PGC-1α function. We conclude that PGC-1α is required for complete skeletal muscle adaptations induced by endurance exercise in mice.

Keywords: endurance exercise, remodeling, signal transduction, gene expression

the understanding of the signaling and molecular mechanisms for endurance exercise-induced skeletal muscle adaptation has been significantly improved in the past decades. Many signal transduction pathways and gene transcriptional machineries have been implicated in skeletal muscle remodeling, including calcineurin (14, 40, 41), Ca2+/calmodulin-dependent protein kinases (CaMK) (50, 54), p38 mitogen-activated protein kinase (MAPK) (3, 49), AMP-activated protein kinase (AMPK) (17, 39, 48, 54), protein kinase D (PKD) (2, 29), myocyte enhancer factor-2 (MEF-2) (35, 51), peroxisome proliferator-activated receptor-γ coactivator-1α (PGC-1α) (23, 33, 44), histone deacetylases (HDAC) (35, 44), and peroxisome proliferator-activated receptor δ (PPARδ) (39, 46). These findings clearly indicate the complexity and the importance of orchestrated signaling cascades and gene regulatory events in skeletal muscle adaptation and laid a solid foundation for a complete elucidation of the regulatory network underlying skeletal muscle plasticity in the future.

Up to date, numerous studies have suggested a pivotal function of PGC-1α in endurance exercise-induced skeletal muscle adaptation. First, both endurance exercise and motor nerve stimulation that mimic the neuromuscular activities in endurance exercise promote Pgc-1α mRNA and protein expression in skeletal muscle (3, 5, 9, 21, 42) as well as PGC-1α activity through posttranslational modification (13). Functionally, forced expression of PGC-1α promotes glycolytic-to-oxidative fiber-type transformation and mitochondrial biogenesis in skeletal muscle (33) with enhanced endurance exercise capacity (12). Consistently, both global and muscle-specific targeted disruptions of the Pgc-1α gene led to a reduced oxidative phenotype in skeletal muscle (23, 32, 34). Muscle-specific deletion of the Pgc-1α gene also resulted in enhanced proinflammatory cytokine expression (24), increased susceptibility to injury (23), and impaired angiogenesis in response to hypoxic stimuli (7). These findings prompted us to employ genetic approach to ascertain the functional role of PGC-1α in endurance exercise-induced skeletal muscle adaptation in a skeletal muscle-specific loss-of-function manner in mice. To this end, we generated skeletal muscle-specific Pgc-1α knockout mice, subjected them and their wild-type littermates to long-term voluntary running, and examined the phenotypic changes in skeletal muscles.

METHODS

Animals.

Mice were housed in temperature-controlled (21°C) quarter with a 12:12-h light-dark cycle with free access to water and normal chow (Purina Chow). Muscle-specific Pgc-1α knockout mice (MKO) were generated by crossbreeding myogenin-Cre mice (generous gift from Dr. Eric Olson) with loxP flanked-Pgc-1α mice (generous gift from Dr. Bruce Spiegelman). Wild-type littermates (WT) were used as control. All animal protocols were approved by the Duke University Institutional Animal Care and Use Committee.

Genotyping.

Mouse DNA was isolated with a phenol-chloroform-based DNA extraction protocol and used in PCR with primers for the Cre allele (5′-AGGTTCGTTCACTCATGGA-3′ and 5′-TCGACCAGTTTAGTTACCC-3′) and the loxP-flanked Pgc-1α allele (5′-TCCAGTAGGCAGAGATTTATGAC-3′ and 5′-TGTCTGGTTTGACAATCTGCTAGGTC-3′). The PCR reaction for the Cre allele was 4 min of initial denaturation at 94°C, 30 cycles of denaturation (94°C for 30 s), annealing (56°C for 30 s) and extension (72°C for 45 s), and a final 7 min of extension at 72°C. The PCR for the loxP-flanked Pgc-1α allele was 4 min of initial denaturation at 94°C, 40 cycles of denaturation (94°C for 30 s), annealing (58°C for 30 s), and extension (72°C for 45 s), and a final 7 min of extension at 72°C. The PCR products were resolved by electrophoresis on 1% agarose gel (containing 1 μg/ml of ethidium bromide), followed by image acquisition under the UV light. Mice with genotype of Pgc-1α +/+:Myog-Cre +/+ or Pgc-1α fl/+:Myog-Cre +/+ will be used as WT mice, and mice with genotype of Pgc-1α fl/fl:Myog-Cre Cre/+ will be used as MKO mice.

Voluntary running.

Both WT and MKO mice (8 wk of age) were housed individually in cages (13 × 13 × 30 cm) that were equipped with running wheels (6, 47). Mice in the exercise group were allowed to run voluntarily for 4 wk while mice in the sedentary group maintained normal cage activity with the running wheels locked. Running activities in the exercise group were recorded and quantified using a desktop PC computer and Dataquest ART Gold Acquisition Software 2.2. After the last bout of running, the running wheels in the exercise groups were locked for 48 h. The mice were humanely killed under fed condition with all the muscle samples harvested within 1 min.

Real-time PCR.

Total RNA was extracted from skeletal muscles using Trizol (Invitrogen) and reverse-transcribed into cDNA as described previously (53). Real-time PCR was performed on an ABI Prism 7000 sequence detection system (Applied Biosystems) with an initial hold temperature of 50°C for 2 min, 95°C for 10 min, followed by cyclic reactions of 95°C for 15 s and 60°C for 1 min. The reagents used in real-time PCR, including primers, probes, and TaqMan PCR Universal master mix (4304437, Applied Biosystems), were purchased from Applied Biosystems: Pgc-1α (Mm00447183_m1), Pgc-1β (Mm00504720_m1), NADH dehydrogenase (ubiquinone) Fe-S protein 1 (Ndufs1) (Mm00523631_m1), estrogen-related receptor α (Esrra) (Mm00433143_m1), and vascular endothelial growth factor (Vegf) (Mm00437306_m1). The DNA sequences of primers and probes were not released to the public due to the proprietary right. The cycle threshold (Ct) was determined with the supplied software, and standard curves were constructed for quantification. 18S rRNA was used as reference to normalize the data.

Immunoblot analysis.

Skeletal muscle sample preparation for immunoblot analysis was described previously (4). The following antibodies were used for immunoblot analysis: cytochrome c oxidase IV (COX IV; ab14744, Abcam Novus Biologicals), cytochrome c (Cyt c; 4272, Cell Signaling Technology), β-actin (sc-81178, Santa Cruz), myoglobin (A0324, DAKO), superoxide dismutase 2 (SOD2; SOD-111, Bioreagent), and myosin heavy chains (MHC) I, IIa, and IIb (BA-F8, SC-71, and BF-F3, respectively, German Collection of Microorganisms and Cell Cultures). Quantification of the proteins was performed using Scion Image software.

Fiber typing.

Skeletal muscle fiber type was determined by immunofluorescence staining for myosin heavy chain proteins in plantaris muscles as described previously (47). After the staining, type I, IIa, IId/x, and IIb fibers were counted for the entire cross section of plantaris muscle and presented as percentage of the total fibers. The percentage area of each fiber type was calculated by Scion Image Software.

Determination of capillary density.

Similar procedures were followed as described previously (47) with the exceptions that different primary and secondary antibodies were used: rat anti-CD31 (MCA1364; Serotec, Raleigh, NC) followed by rhodamine red-X-conjugated goat anti-rat IgG (The Jackson Laboratory). Total number of capillaries and the surface area of the entire cross section for each muscle were measured with Scion Image software (Scion, Frederick, MD) and presented as capillary density (capillaries/mm2).

Transmission electron microscopy.

Transmission electron microscopy analysis was performed at Duke Electron Microscopy Service for plantaris muscle longitudinal sections from sedentary and exercised WT and MKO mice according to the method described previously (36). Defective mitochondria out of total mitochondria were counted in ≥14 randomly acquired TEM images for each group of mice.

Statistical analysis.

The results are presented as means ± SE. Comparisons among different treatments were analyzed for statistical significance using two-way ANOVA followed by the Student-Newman-Keuls test with P < 0.05 as statistically significant.

RESULTS

Pgc-1α MKO mice have normal IIb-to-IIa fiber-type transformation in response to voluntary running.

To determine if PGC-1α is essential for endurance exercise-induced skeletal muscle adaptation, we obtained MKO and WT mice by crossbreeding (Fig. 1A). The efficacy of the gene deletion was confirmed by real-time PCR analysis of total RNA from plantaris muscles, showing that MKO mice only had trace amount of Pgc-1α mRNA expression (Fig. 1B). Similar to the findings from previous studies (23, 24, 32), MKO mice showed a trend of decreased expression of Esrra, Ndufs1, and Vegf mRNA with no appreciable changes in Pgc-1β mRNA in skeletal muscle (Fig. 1B). When we subjected MKO and WT mice (male, 8 wk of age) to 4 wk of voluntary running, the daily running distance for each group increased gradually, reaching a steady-state level after 2 wk (Fig. 1C). No significant differences in running distance were observed between MKO and WT mice (Fig. 1D). Both MKO and WT mice had moderate but significant increases in the heart weight (normalized by body weight) following endurance exercise training (Fig. 1E), indicating similar degrees of endurance exercise-induced cardiac hypertrophy. Neither the body weight [24.8 ± 0.5, 23.7 ± 0.9, 24.6 ± 1.4, and 23.2 ± 0.4 g for the WT sedentary (WT-Sed), WT endurance exercise-trained (WT-Ex), MKO-Sed, and MKO-Ex groups, respectively, P > 0.05] nor plantaris muscle weight (0.62 ± 0.01, 0.59 ± 0.02, 0.59 ± 0.02, and 0.61 ± 0.01 mg/g body wt for WT-Sed, WT-Ex, MKO-Sed and MKO-Ex groups, respectively, P > 0.05) showed any significant difference among different groups.

An external file that holds a picture, illustration, etc.
Object name is zh00031061950001.jpg

Voluntary exercise training in muscle-specific peroxisome proliferator-activated receptor γ coactivator-1α gene (Pgc-1α) knockout (MKO) and wild-type (WT) mice. A: gel images of PCR products amplified from genomic DNA from WT (Pgc-1α +/+:Myog-Cre +/+ or Pgc-1α fl/+:Myog-Cre +/+) and MKO (Pgc-1α fl/fl:Myog-Cre Cre/+) mice. B: total RNA from plantaris muscles of WT and MKO mice were assayed for mRNA expression of Pgc-1α, Pgc-1β, estrogen-related receptor α (Esrra), NADH dehydrogenase (ubiquinone) Fe-S protein 1 (Ndufs1), and vascular endothelial growth factor (Vegf) by real-time PCR. 18S ribosomal RNA was used to normalize the expression (n = 6). *P < 0.05, ***P < 0.001. C: representative running activity recording. Six continuous voluntary running activity recordings are shown for WT and MKO mice during the 4 wk of voluntary running. Each of the clusters of running activities (spikes) that correspond to a duration of half-day represents the nocturnal running activity at night. D: daily voluntary running distance in WT and MKO mice during the 4 wk of voluntary running (P = 0.99). Values are means ± SE (n = 7). E: heart weight (normalized by body weight) in sedentary (Sed) and 4-wk endurance exercise-trained (Ex) WT and MKO mice (n = 6–7). *P < 0.05.

Since plantaris muscles are actively recruited during voluntary running (51), and fiber-type analysis could be done precisely for the cross section in the middle portion of the muscle, we performed immunofluorescence-based fiber-type analysis in plantaris muscles following 4 wk of voluntary running as described (4). Our results showed that endurance exercise induced a significant increase in the percentage of type IIa myofibers in WT mice with a concurrent decrease in the percentage of type IIb fibers (Fig. 2A and Table 1). The percentage of type I fiber is extremely low in plantaris muscles, and endurance exercise training with the duration had no effect on the percentage of type I fibers in either genotypes. The increases in the percentage of type IIa myofibers in plantaris muscles were similar between MKO and WT mice, indicating this fiber-type transformation was independent of the genotypes (Fig. 2A and Table 1). To further confirm this result, we performed immunoblot analysis to quantify the expression of different isoforms of MHC proteins. As shown in Fig. 2, B and C, voluntary running induced significant increases in MHC IIa protein expression in plantaris muscles with a trend of reduction of MHC IIb protein; these changes were not affected by muscle-specific disruption of the Pgc-1α gene. Consistent with the immunofluorescence-based fiber type analysis, we could only detect trace amount of MHC I protein expression, and it was not affected by voluntary running. We observed the same MHC IIa protein induction by immunoblot analysis in two other predominantly glycolytic muscles, gastrocnemius (GA) and extensor digitorum longus (EDL), independent of the genotype (see Supplementary Figure S1, available with the online version of this article).

An external file that holds a picture, illustration, etc.
Object name is zh00031061950002.jpg

Immunofluorescence and immunoblot analyses for fiber-type composition in plantaris muscles of sedentary (Sed) and 4-wk endurance exercise-trained (Ex) WT and MKO mice. A: cross sections of plantaris muscles were immunostained for myosin heavy chains (MHC) I (red), IIa (blue), and IIb (green). Bars, 500 μm. All the fibers were counted and measured for cross-sectional area for calculation of fiber-type composition (please see Table 1). B: immunoblot analysis of MHC I, MHC IIa, and MHC IIb in plantaris muscles from sedentary (S) and endurance exercise-trained (E) mice with β-actin as loading control. Soleus muscle (Sol) was used as a positive control for MHC I. C: quantifications of MHC I, MHC IIa, and MHC IIb protein expression in plantaris muscles. Values are means ± SE (n = 6–7). **P < 0.01.

Table 1.

Comparison of fiber type composition in plantaris muscles in response to voluntary running in MKO and WT mice

WT
MKO
Sed (n = 6)Ex (n = 5)Sed (n = 6)Ex (n = 7)
Fiber-type composition
MHC I
    %Total fibers1.42 ± 0.491.49 ± 0.593.32 ± 0.733.51 ± 0.65
    %Area0.46 ± 0.170.92 ± 0.541.50 ± 0.351.39 ± 0.27
MHC IIa
    %Total fibers21.1 ± 1.129.5 ± 2.522.5 ± 1.129.6 ± 1.3
    %Area15.1 ± 1.626.3 ± 5.0*20.1 ± 1.032.4 ± 2.5
MHC IId/x
    %Total fibers29.5 ± 1.330.0 ± 2.627.7 ± 1.126.4 ± 1.7
    %Area20.7 ± 1.226.8 ± 5.2*24.9 ± 1.328.6 ± 2.2
MHC IIb
    %Total fibers48.0 ± 1.639.0 ± 4.4*46.5 ± 1.540.5 ± 2.2
    %Area63.7 ± 2.746.0 ± 9.6*53.5 ± 2.037.6 ± 3.9*

Values are means ± SE. MKO, muscle-specific Pgc-1α gene knockout mice; WT, wild-type littermates; Sed, sedentary; Ex, endurance exercise trained; MHC, myosin heavy chain; %Total fibers, fiber type as percentage of total fibers counted; %Area, fiber type area as percentage of total area.

*P < 0.05,
P < 0.01 vs. Sed group of the same genotype.
P < 0.05 vs. WT group of the same treatment.

Muscle-specific deletion of the Pgc-1α gene attenuates endurance exercise-induced mitochondrial biogenesis.

PGC-1α is a pivotal regulator of mitochondrial biogenesis by interacting with and stimulating the expression of transcription factors that function in control of nuclear-mitochondrial interactions, such as nuclear respiratory factor 1 (NRF1) and NRF2 (52). Despite the fact that overexpression of PGC-1α in cultured muscle cells and adult skeletal muscle induces mitochondrial biogenesis (33, 52), it was not known whether PGC-1α is required for endurance exercise-induced mitochondrial biogenesis, particularly in light of the recent finding that some exercise-responsive genes, including genes that function in mitochondrial oxidative phosphorylation, were not affected in mice with global knockout of the Pgc-1α gene (31). To address this important question, we performed immunoblot analysis for COX IV and Cyt c in skeletal muscles from MKO and WT mice following 4 wk of endurance exercise. These two proteins are critical components of the mitochondrial electron transport chain, and their expression levels are indicative of mitochondrial biogenesis (15, 26, 45). Following 4 wk of voluntary running, WT mice showed a 1.8-fold increase in COX IV expression (P < 0.001) while MKO mice showed a significantly less increase (1.3-fold, P < 0.05) (Fig. 3, A and B). The posttraining level of COX IV protein in MKO was significantly lower than that in WT mice (P < 0.001). Similar results were obtained for Cyt c protein; endurance exercise training induced a significant (1.6-fold) increase of Cyt c protein in WT mice (P < 0.001), but not in MKO mice (1.3-fold; P = 0.10) (Fig. 3, A and B). Interestingly, endurance exercise-induced expression of myoglobin, which is known to be upregulated in cultured muscle cells and in skeletal muscle in transgenic mice where the Pgc-1α gene is overexpressed (25, 33), was not affected by muscle-specific deletion of the Pgc-1α gene (P < 0.05 and P < 0.01 for WT and MKO mice, respectively). Similar observations for these proteins were made in three other glycolytic muscles, GA, EDL, and white vastus lateralis (WV) muscles (see Supplementary Figure S1, available with the online version of this article).

An external file that holds a picture, illustration, etc.
Object name is zh00031061950003.jpg

Immunoblot analysis for myoglobin and mitochondrial enzymes and immunofluorescence analysis for capillary density in skeletal muscles of sedentary (Sed) and 4-wk endurance exercise-trained (Ex) WT and MKO mice. A: immunoblot analysis of cytochrome oxidase IV (COX IV), cytochrome c (Cyt c), and myoglobin (Mb) protein expression in plantaris muscles from sedentary (S) and endurance exercise trained mice (E) with β-actin as loading control. Each lane was loaded with 40 μg of muscle protein from whole muscle homogenates. B: quantifications of COX IV, Cyt c and Mb protein expression in plantaris muscle (n = 6–7). *P < 0.05, **P < 0.01, ***P < 0.001. C: immunofluorescence staining of endothelial cells with anti-CD31 antibody in cross sections of plantaris muscle. Bars = 100 μm. D: quantifications of capillary density in plantaris muscle (n = 6–7). *P < 0.05, ***P < 0.001, respectively.

Muscle-specific deletion of the Pgc-1α gene attenuates endurance exercise-induced angiogenesis.

Angiogenesis is one of the important adaptations of skeletal muscle in response to endurance exercise training, and PGC-1α has recently been shown to play an important role in hypoxia-induced VEGF expression and angiogenesis independent of hypoxia-inducible factor-1 (HIF-1) in skeletal muscles (7). Considering this and the fact that MKO mice present a trend for reduced Vegf mRNA expression (Fig. 1B), we hypothesized that PGC-1α plays an essential role in endurance exercise-induced angiogenesis and performed immunofluorescence by using specific antibodies against platelet endothelial cell adhesion molecule-1 (PECAM-1, CD31) to evaluate capillary density in skeletal muscle following voluntary running. Consistent with our hypothesis, endurance exercise induced a ~60% increase in the number of CD31-positive endothelial cells per unit cross-sectional area (capillary density) in the WT mice. This adaptation was abolished in plantaris muscles from mice with muscle-specific deletion of the Pgc-1α gene (P = 0.17) (Fig. 3, C and D).

Voluntary running corrects abnormal mitochondrial morphology and restores SOD2 expression in skeletal muscles in MKO mice.

It has been shown previously that muscle-specific deletion of the Pgc-1α gene leads to enhanced expression of IL-6 and oxidative genes and increased susceptibility to muscle injury (24). We performed electron microscopy for plantaris muscle sections and found that MKO mice had significantly more vacuolated mitochondria (18.6 ± 3.2%) in both glycolytic and oxidative myofibers than WT mice (10.2 ± 3.2%). The vacuolated mitochondria showed reduced density of the matrix, disarray of the cristae, and significant enlargement (Fig. 4, A and B). Importantly, 4 wk of voluntary running resulted in significant reduction in the number and severity of mitochondrial vacuolation in both WT (4.8 ± 0.7%, P < 0.05 vs. WT-Sed mice) and MKO mice (3.9 ± 0.9%, P < 0.01 vs. MKO-Sed mice) (Fig. 4, A and B). Immunoblot analysis revealed that MKO mice had significantly less SOD2 protein expression in plantaris muscles compared with the wild-type littermates, which could be completely corrected by voluntary running (Fig. 4, C and D). These findings suggest that deletion of the Pgc-1α gene resulted in significantly compromised mitochondria integrity and impaired mitochondrial SOD2 expression in skeletal muscle, which could be restored by endurance exercise training in the absence of functional PGC-1α.

An external file that holds a picture, illustration, etc.
Object name is zh00031061950004.jpg

Transmission electron microscopy (TEM) analysis for mitochondrial morphology and immunoblot analysis for antioxidant enzyme expression in plantaris muscles of sedentary (Sed) and 4-wk endurance exercise-trained (Ex) WT and MKO mice. A: TEM images of longitudinal plantaris muscle sections of glycolytic and oxidative myofibers. Arrows show degenerative mitochondrion. To show detailed structure of some degenerative mitochondria, cropped images are shown with 2.5× magnification of the original images on one corner. Bars = 1 μm. B: quantification of abnormal mitochondria in ≥14 randomly acquired images as a percentage of total mitochondria. *P < 0.05, ***P < 0.001, respectively. C: immunoblot analysis of superoxide dismutase 2 (SOD2) protein expression in plantaris muscles from sedentary (S) and endurance exercise trained mice (E) with β-actin as loading control. Each lane was loaded with 40 μg of muscle protein from whole muscle homogenates. D: quantifications of SOD2 protein expression in plantaris muscle (n = 5–7). *P < 0.05.

DISCUSSION

One of the important questions in exercise physiology is about the regulatory factor(s) that is functionally responsible for endurance exercise-induced skeletal muscle adaptation. Since the cloning of the Pgc-1α gene, there have been increasing interests in its function in skeletal muscle plasticity. The most intriguing finding is the formation of significantly more oxidative myofibers when the Pgc-1α gene is overexpressed in skeletal muscle in transgenic mice (33). However, it was not known whether PGC-1α was required for endurance exercise-induced skeletal muscle adaptation and what aspects, if not all, of muscle adaptation would be affected when the Pgc-1α gene was deleted. This study confirmed the essential function of PGC-1α in endurance exercise-induced adaptation based on the findings that MKO mice had significant attenuation of 1) the induction of mitochondrial electron transport protein expression, and 2) new capillary formation in skeletal muscles. These two adaptations constitute the major adaptive processes that confer the oxidative metabolic phenotype leading to improved oxidative phosphorylation and microcirculation during exercise.

The voluntary running model has been used to induce skeletal muscle adaptation in rodents (6). While the periodic bursts of running activity are not the same as the endurance exercise in humans, the phenotypic adaptations are quite similar (4, 6, 47). The MKO mice were able to run similar distance as WT mice with a nearly twofold increase in daily running distance after 2 wk of endurance exercise. Therefore, the attenuated muscle adaptation in mitochondrial biogenesis and angiogenesis in MKO mice is clearly not due to reduced running activity in these mice. MKO had a similar degree of physiological cardiac hypertrophy (heart weight normalized by body weight) compared with WT mice, further supporting this notion. Furthermore, induced expression of other exercise responsive genes, such as myoglobin and MHC IIa, was not affected by the Pgc-1α gene deletion in skeletal muscle, providing biochemical evidence for the exercise perturbation.

The similar daily running distance between MKO and WT mice seems to be contradictory to the substantial attenuation of endurance exercise-induced COX IV and Cyt c expression as these proteins play important role in mitochondrial oxidative phosphorylation and hence endurance exercise capacity. There are at least three possible reasons for this apparently paradoxical finding. First and most likely, mitochondrial function and/or efficiency may be enhanced independent of mitochondrial biogenesis. In other words, in the absence of increased mitochondrial electron transport protein expression, mitochondrial function may still be improved through other processes, such as posttranslational modification and/or improved physical structure of the electron transport chain complexes. Our finding that MKO mice had a significant amount of abnormal mitochondria and reduced SOD2 protein expression that could be corrected by voluntary running support to a certain degree the independence of the mitochondrial remodeling process on PGC-1α function. Consistent with this notion, Adhihetty et al. (1) have recently shown that Pgc-1α gene disruption impaired mitochondrial function (state 3 and state 4 respiration) in isolated mitochondria from skeletal muscle of sedentary animals, but that endurance exercise training led to enhanced mitochondrial function regardless of the genetic background. Second, other muscle adaptive processes, such as those related to calcium handling and the contractile apparatus, may result in compensation for the Pgc-1α gene deletion-induced loss of oxidative function in skeletal muscles. Finally, although voluntary running is a great physiological exercise model to induce skeletal muscle adaptation, the daily running distance on the voluntary wheel is not an index of endurance capacity. It would be very interesting to determine if MKO mice have lower endurance capacity compared with WT mice and if endurance exercise training improves exercise capacity in these mice.

Mitochondrial biogenesis, the process by which new mitochondria are formed in the cell along with increased mitochondrial enzyme expression/activity, is not only a hallmark of skeletal muscle adaptation (20, 27, 28) but is also pivotal for the enhanced exercise capacity (10, 27). Much of the recent attentions have been paid to PGC-1α function in endurance exercise-induced mitochondrial biogenesis since activation of Pgc-1α transcription (9, 42) and activity (13) occurs upon a single bout of endurance exercise, and transgenic overexpression of the Pgc-1α gene or upstream signaling molecules in skeletal muscle resulted in enhanced mitochondrial biogenesis (3, 19, 33). While we were conducting this study in muscle-specific Pgc-1α knockout mice, Leick et al. (31) employed a mouse model of global gene disruption of the Pgc-1α gene and obtained surprising results indicating that Pgc-1α is not mandatory for exercise- and training-induced adaptive gene responses. Since the mice with the global disruption of the Pgc-1α gene had hyperactivity due to lesions in the central nervous system, an approach of skeletal muscle-specific gene disruption is desirable to minimize the potential confounding factors. Another note to consider is that two of the protein markers, hexokinase II (HK II) and aminolevulinate synthase 1 (ALAS1) used in the aforementioned study are not mitochondrial enzymes, and the other one, COX I, is a protein that has been shown not to be controlled by PGC-1α (18). Our finding that endurance exercise-induced COX IV and Cyt c protein expression were significantly attenuated in MKO mice is consistent with an essential functional role of PGC-1α in mediating endurance exercise-induced mitochondrial biogenesis. A final note on mitochondrial adaptations in MKO is the observation that mitochondrial morphology and SOD2 protein expression were improved with endurance exercise training in these animals. Thus, whereas normal PGC-1α expression seems to be required for mitochondrial maintenance in sedentary animals, the mechanisms involved in endurance exercise-induced improvement in mitochondrial morphology remain to be determined.

Angiogenesis, the formation of new capillary from the existing vasculature, is equally important to improved functional capacity after exercise training (11, 22). It has even been postulated that angiogenesis precedes and play a permissive role for the other adaptive processes, such as fiber-type transformation (47). The recent study by Arany et al. (7) showed that PGC-1α is necessary and sufficient for hypoxia-induced angiogenesis in skeletal muscle. This raised the possibility of a functional role of PGC-1α in endurance exercise-induced angiogenesis although it is still debatable whether hypoxia constitutes an endurance exercise signal for phenotypic adaptations. Interestingly, mice with global gene disruption of the Pgc-1α gene showed impaired VEGF and CD31 expression in skeletal muscle in response to endurance exercise training (30) despite the findings that these mice had relatively normal exercise-induced gene responses (31). Our findings that MKO mice had abolished increase of capillary density in skeletal muscle following 4 wk of voluntary running provide direct evidence that PGC-1α is essential for endurance exercise-induced angiogenesis. In light of the findings that hypoxia-induced angiogenesis in skeletal muscle is mediated by PGC-1α independent of HIF-1, it would be interesting to determine if endurance exercise-induced angiogenesis in skeletal muscle is achieved through the same mechanism.

In addition to the phenotypic changes that confer characteristics in metabolism, endurance exercise also promotes IIb-to-IIa fiber-type transformation (16, 47). While overexpression of Pgc-1α in skeletal muscle promotes more oxidative myofiber formation (33), we showed here that voluntary exercise-induced IIb-to-IIa transformation was not affected by the deletion of the Pgc-1α gene in skeletal muscle, suggesting that endurance exercise-induced fiber-type transformation in skeletal muscle could occur independent of the function of the Pgc-1α gene. These findings have several implications. First, considering the impaired angiogenesis in the muscle-specific knockout mice, one might argue that angiogenesis is not obligatory for endurance exercise-induced fiber-type switching. In other words, these findings are not consistent with the hypothesis that angiogenesis plays an instrumental role for fiber-type transformation. Second, it has previously been shown that overexpression of active upstream signaling molecules in skeletal muscle that upregulate PGC-1α expression can promote mitochondrial biogenesis without inducing shifts in fiber-type composition (3, 19). We have also obtained new evidence in skeletal muscle-specific p38 knockout mice that endurance exercise-induced mitochondrial biogenesis and angiogenesis in skeletal muscle, but not fiber-type transformation, are dependent on the upstream p38γ MAPK signaling pathway (43). The findings in this study once again in a tissue-specific loss-of-function mouse model genetically separate the endurance exercise-induced mitochondrial biogenesis and angiogenesis from the contractile adaptation, specifically IIb-to-IIa fiber-type transformation. All these findings collectively support the view that PGC-1α and its regulatory control axis play a functional role in complete skeletal muscle adaptations that lead to improved microcirculatory function and oxidative metabolism, but their function seems not absolutely required for endurance exercise-induced contractile adaptation in mice.

It should be pointed out that although we have employed skeletal muscle-specific Pgc-1α knockout mice to minimize the possibility of developmental and interorgan influences from the disruption of the gene globally, we could not rule out the possibility of other regulatory factors that function in a compensatory manner. For example, Pgc-1β has recently been shown to promote formation of type IIx fibers (8), although the decreased Pgc-1β expression in human skeletal muscle by endurance exercise training (38) appears to be contradictory to this potential function in promoting fiber-type transformation from IIb to IIa. The possibility that a redundant functional role of a different isoform of PGC-1α as a result of alternative promoter (37) in MKO mice could be ruled out as we deleted exons 3, 4, and 5 of the Pgc-1α gene and prevented the formation of this PGC-1α isoform. The fact that endurance exercise-induced mitochondrial biogenesis and angiogenesis are significantly attenuated in Pgc-1α MKO mice suggests that the essential function of PGC-1α in these fundamental adaptive processes could not be compensated adequately by redundant systems if they existed.

In summary, our studies revealed that Pgc-1α plays an important functional role for endurance exercise-induced mitochondrial biogenesis and angiogenesis but not for IIb-to-IIa fiber-type transformation in mouse skeletal muscle. Our findings provide evidence in a loss-of-function mode that some of the metabolic adaptations could be genetically separated from the contractile adaptation. The improved understanding of the molecular mechanisms by which endurance exercise training promotes skeletal muscle adaptation will be instrumental for future development of novel interventions for many chronic diseases that have been shown to be benefited from regular exercise.

DISCLOSURES

No conflicts of interest are declared by the authors.

Supplementary Material

[Supplemental Figure]

REFERENCES

1. Adhihetty PJ, Uguccioni G, Leick L, Hidalgo J, Pilegaard H, Hood DA. The role of PGC-1α on mitochondrial function and apoptotic susceptibility in muscle. Am J Physiol Cell Physiol 297: C217–C225, 2009 [Abstract] [Google Scholar]
2. Akimoto T, Li P, Yan Z. Functional interaction of regulatory factors with the Pgc-1α promoter in response to exercise by in vivo imaging. Am J Physiol Cell Physiol 295: C288–C292, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
3. Akimoto T, Pohnert SC, Li P, Zhang M, Gumbs C, Rosenberg PB, Williams RS, Yan Z. Exercise stimulates Pgc-1alpha transcription in skeletal muscle through activation of the p38 MAPK pathway. J Biol Chem 280: 19587–19593, 2005 [Abstract] [Google Scholar]
4. Akimoto T, Ribar TJ, Williams RS, Yan Z. Skeletal muscle adaptation in response to voluntary running in Ca2+/calmodulin-dependent protein kinase IV-deficient mice. Am J Physiol Cell Physiol 287: C1311–C1319, 2004 [Abstract] [Google Scholar]
5. Akimoto T, Sorg BS, Yan Z. Real-time imaging of peroxisome proliferator-activated receptor-gamma coactivator-1α promoter activity in skeletal muscles of living mice. Am J Physiol Cell Physiol 287: C790–C796, 2004 [Abstract] [Google Scholar]
6. Allen DL, Harrison BC, Maass A, Bell ML, Byrnes WC, Leinwand LA. Cardiac and skeletal muscle adaptations to voluntary wheel running in the mouse. J Appl Physiol 90: 1900–1908, 2001 [Abstract] [Google Scholar]
7. Arany Z, Foo SY, Ma Y, Ruas JL, Bommi-Reddy A, Girnun G, Cooper M, Laznik D, Chinsomboon J, Rangwala SM, Baek KH, Rosenzweig A, Spiegelman BM. HIF-independent regulation of VEGF and angiogenesis by the transcriptional coactivator PGC-1alpha. Nature 451: 1008–1012, 2008 [Abstract] [Google Scholar]
8. Arany Z, Lebrasseur N, Morris C, Smith E, Yang W, Ma Y, Chin S, Spiegelman BM. The transcriptional coactivator PGC-1beta drives the formation of oxidative type IIX fibers in skeletal muscle. Cell Metab 5: 35–46, 2007 [Abstract] [Google Scholar]
9. Baar K, Wende AR, Jones TE, Marison M, Nolte LA, Chen M, Kelly DP, Holloszy JO. Adaptations of skeletal muscle to exercise: rapid increase in the transcriptional coactivator PGC-1. FASEB J 16: 1879–1886, 2002 [Abstract] [Google Scholar]
10. Baldwin KM, Klinkerfuss GH, Terjung RL, Mole PA, Holloszy JO. Respiratory capacity of white, red, and intermediate muscle: adaptative response to exercise. Am J Physiol 222: 373–378, 1972 [Abstract] [Google Scholar]
11. Breen EC, Johnson EC, Wagner H, Tseng HM, Sung LA, Wagner PD. Angiogenic growth factor mRNA responses in muscle to a single bout of exercise. J Appl Physiol 81: 355–361, 1996 [Abstract] [Google Scholar]
12. Calvo JA, Daniels TG, Wang X, Paul A, Lin J, Spiegelman BM, Stevenson SC, Rangwala SM. Muscle-specific expression of PPARγ coactivator-1α improves exercise performance and increases peak oxygen uptake. J Appl Physiol 104: 1304–1312, 2008 [Abstract] [Google Scholar]
13. Canto C, Gerhart-Hines Z, Feige JN, Lagouge M, Noriega L, Milne JC, Elliott PJ, Puigserver P, Auwerx J. AMPK regulates energy expenditure by modulating NAD+ metabolism and SIRT1 activity. Nature 458: 1056–1060, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
14. Chin ER, Olson EN, Richardson JA, Yang Q, Humphries C, Shelton JM, Wu H, Zhu W, Bassel-Duby R, Williams RS. A calcineurin-dependent transcriptional pathway controls skeletal muscle fiber type. Genes Dev 12: 2499–2509, 1998 [Europe PMC free article] [Abstract] [Google Scholar]
15. Dudley GA, Abraham WM, Terjung RL. Influence of exercise intensity and duration on biochemical adaptations in skeletal muscle. J Appl Physiol 53: 844–850, 1982 [Abstract] [Google Scholar]
16. Fitzsimons DP, Diffee GM, Herrick RE, Baldwin KM. Effects of endurance exercise on isomyosin patterns in fast- and slow-twitch skeletal muscles. J Appl Physiol 68: 1950–1955, 1990 [Abstract] [Google Scholar]
17. Fujii N, Hayashi T, Hirshman MF, Smith JT, Habinowski SA, Kaijser L, Mu J, Ljungqvist O, Birnbaum MJ, Witters LA, Thorell A, Goodyear LJ. Exercise induces isoform-specific increase in 5′-AMP-activated protein kinase activity in human skeletal muscle. Biochem Biophys Res Commun 273: 1150–1155, 2000 [Abstract] [Google Scholar]
18. Garcia-Roves PM, Huss J, Holloszy JO. Role of calcineurin in exercise-induced mitochondrial biogenesis. Am J Physiol Endocrinol Metab 290: E1172–E1179, 2006 [Abstract] [Google Scholar]
19. Garcia-Roves PM, Osler ME, Holmstrom MH, Zierath JR. Gain-of-function R225Q mutation in AMP-activated protein kinase gamma3 subunit increases mitochondrial biogenesis in glycolytic skeletal muscle. J Biol Chem 283: 35724–35734, 2008 [Abstract] [Google Scholar]
20. Gollnick PD, King DW. Effect of exercise and training on mitochondria of rat skeletal muscle. Am J Physiol 216: 1502–1509, 1969 [Abstract] [Google Scholar]
21. Goto M, Terada S, Kato M, Katoh M, Yokozeki T, Tabata I, Shimokawa T. cDNA Cloning and mRNA analysis of PGC-1 in epitrochlearis muscle in swimming-exercised rats. Biochem Biophys Res Commun 274: 350–354, 2000 [Abstract] [Google Scholar]
22. Gute D, Laughlin MH, Amann JF. Regional changes in capillary supply in skeletal muscle of interval-sprint and low-intensity, endurance-trained rats. Microcirculation 1: 183–193, 1994 [Abstract] [Google Scholar]
23. Handschin C, Chin S, Li P, Liu F, Maratos-Flier E, Lebrasseur NK, Yan Z, Spiegelman BM. Skeletal muscle fiber-type switching, exercise intolerance, and myopathy in PGC-1alpha muscle-specific knock-out animals. J Biol Chem 282: 30014–30021, 2007 [Abstract] [Google Scholar]
24. Handschin C, Choi CS, Chin S, Kim S, Kawamori D, Kurpad AJ, Neubauer N, Hu J, Mootha VK, Kim YB, Kulkarni RN, Shulman GI, Spiegelman BM. Abnormal glucose homeostasis in skeletal muscle-specific PGC-1alpha knockout mice reveals skeletal muscle-pancreatic beta cell crosstalk. J Clin Invest 117: 3463–3474, 2007 [Europe PMC free article] [Abstract] [Google Scholar]
25. Handschin C, Rhee J, Lin J, Tarr PT, Spiegelman BM. An autoregulatory loop controls peroxisome proliferator-activated receptor gamma coactivator 1alpha expression in muscle. Proc Natl Acad Sci USA 100: 7111–7116, 2003 [Europe PMC free article] [Abstract] [Google Scholar]
26. Hickson RC. Skeletal muscle cytochrome c and myoglobin, endurance, and frequency of training. J Appl Physiol 51: 746–749, 1981 [Abstract] [Google Scholar]
27. Holloszy JO. Biochemical adaptations in muscle. Effects of exercise on mitochondrial oxygen uptake and respiratory enzyme activity in skeletal muscle. J Biol Chem 242: 2278–2282, 1967 [Abstract] [Google Scholar]
28. Hoppeler H, Luthi P, Claassen H, Weibel ER, Howald H. The ultrastructure of the normal human skeletal muscle. A morphometric analysis on untrained men, women and well-trained orienteers. Pflügers Arch 344: 217–232, 1973 [Abstract] [Google Scholar]
29. Kim MS, Fielitz J, McAnally J, Shelton JM, Lemon DD, McKinsey TA, Richardson JA, Bassel-Duby R, Olson EN. Protein kinase D1 stimulates MEF2 activity in skeletal muscle and enhances muscle performance. Mol Cell Biol 28: 3600–3609, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
30. Leick L, Hellsten Y, Fentz J, Lyngby SS, Wojtaszewski JF, Hidalgo J, Pilegaard H. PGC-1alpha mediates exercise-induced skeletal muscle VEGF expression in mice. Am J Physiol Endocrinol Metab 297: E92–E103, 2009 [Abstract] [Google Scholar]
31. Leick L, Wojtaszewski JF, Johansen ST, Kiilerich K, Comes G, Hellsten Y, Hidalgo J, Pilegaard H. PGC-1α is not mandatory for exercise- and training-induced adaptive gene responses in mouse skeletal muscle. Am J Physiol Endocrinol Metab 294: E463–E474, 2008 [Abstract] [Google Scholar]
32. Leone TC, Lehman JJ, Finck BN, Schaeffer PJ, Wende AR, Boudina S, Courtois M, Wozniak DF, Sambandam N, Bernal-Mizrachi C, Chen Z, Holloszy JO, Medeiros DM, Schmidt RE, Saffitz JE, Abel ED, Semenkovich CF, Kelly DP. PGC-1alpha deficiency causes multi-system energy metabolic derangements: muscle dysfunction, abnormal weight control and hepatic steatosis. PLoS Biol 3: e101, 2005 [Europe PMC free article] [Abstract] [Google Scholar]
33. Lin J, Wu H, Tarr PT, Zhang CY, Wu Z, Boss O, Michael LF, Puigserver P, Isotani E, Olson EN, Lowell BB, Bassel-Duby R, Spiegelman BM. Transcriptional co-activator PGC-1 alpha drives the formation of slow-twitch muscle fibres. Nature 418: 797–801, 2002 [Abstract] [Google Scholar]
34. Lin J, Wu PH, Tarr PT, Lindenberg KS, St-Pierre J, Zhang CY, Mootha VK, Jager S, Vianna CR, Reznick RM, Cui L, Manieri M, Donovan MX, Wu Z, Cooper MP, Fan MC, Rohas LM, Zavacki AM, Cinti S, Shulman GI, Lowell BB, Krainc D, Spiegelman BM. Defects in adaptive energy metabolism with CNS-linked hyperactivity in PGC-1alpha null mice. Cell 119: 121–135, 2004 [Abstract] [Google Scholar]
35. McGee SL, Hargreaves M. Exercise and myocyte enhancer factor 2 regulation in human skeletal muscle. Diabetes 53: 1208–1214, 2004 [Abstract] [Google Scholar]
36. Miller SE, Howell DN. Viral infections in the acquired immunodeficiency syndrome. J Electron Microsc Tech 8: 41–78, 1988 [Europe PMC free article] [Abstract] [Google Scholar]
37. Miura S, Kai Y, Kamei Y, Ezaki O. Isoform-specific increases in murine skeletal muscle peroxisome proliferator-activated receptor-gamma coactivator-1alpha (PGC-1alpha) mRNA in response to beta2-adrenergic receptor activation and exercise. Endocrinology 149: 4527–4533, 2008 [Abstract] [Google Scholar]
38. Mortensen OH, Plomgaard P, Fischer CP, Hansen AK, Pilegaard H, Pedersen BK. PGC-1beta is downregulated by training in human skeletal muscle: no effect of training twice every second day vs. once daily on expression of the PGC-1 family. J Appl Physiol 103: 1536–1542, 2007 [Abstract] [Google Scholar]
39. Narkar VA, Downes M, Yu RT, Embler E, Wang YX, Banayo E, Mihaylova MM, Nelson MC, Zou Y, Juguilon H, Kang H, Shaw RJ, Evans RM. AMPK and PPARdelta agonists are exercise mimetics. Cell 134: 405–415, 2008 [Europe PMC free article] [Abstract] [Google Scholar]
40. Naya FJ, Mercer B, Shelton J, Richardson JA, Williams RS, Olson EN. Stimulation of slow skeletal muscle fiber gene expression by calcineurin in vivo. J Biol Chem 275: 4545–4548, 2000 [Abstract] [Google Scholar]
41. Parsons SA, Wilkins BJ, Bueno OF, Molkentin JD. Altered skeletal muscle phenotypes in calcineurin Aalpha and Abeta gene-targeted mice. Mol Cell Biol 23: 4331–4343, 2003 [Europe PMC free article] [Abstract] [Google Scholar]
42. Pilegaard H, Saltin B, Neufer PD. Exercise induces transient transcriptional activation of the PGC-1alpha gene in human skeletal muscle. J Physiol 546: 851–858, 2003 [Abstract] [Google Scholar]
43. Pogozelski AR, Geng T, Li P, Yin X, Lira VA, Zhang M, Chi JT, Yan Z. p38gamma Mitogen-activated protein kinase is a key regulator in skeletal muscle metabolic adaptation in mice. PLoS One 4: e7934, 2009 [Europe PMC free article] [Abstract] [Google Scholar]
44. Potthoff MJ, Wu H, Arnold MA, Shelton JM, Backs J, McAnally J, Richardson JA, Bassel-Duby R, Olson EN. Histone deacetylase degradation and MEF2 activation promote the formation of slow-twitch myofibers. J Clin Invest 117: 2459–2467, 2007 [Europe PMC free article] [Abstract] [Google Scholar]
45. Steenaart NA, Shore GC. Mitochondrial cytochrome c oxidase subunit IV is phosphorylated by an endogenous kinase. FEBS Lett 415: 294–298, 1997 [Abstract] [Google Scholar]
46. Wang YX, Zhang CL, Yu RT, Cho HK, Nelson MC, Bayuga-Ocampo CR, Ham J, Kang H, Evans RM. Regulation of muscle fiber type and running endurance by PPARdelta. PLoS Biol 2: e294, 2004 [Europe PMC free article] [Abstract] [Google Scholar]
47. Waters RE, Rotevatn S, Li P, Annex BH, Yan Z. Voluntary running induces fiber type-specific angiogenesis in mouse skeletal muscle. Am J Physiol Cell Physiol 287: C1342–C1348, 2004 [Abstract] [Google Scholar]
48. Wojtaszewski JF, Nielsen P, Hansen BF, Richter EA, Kiens B. Isoform-specific and exercise intensity-dependent activation of 5′-AMP-activated protein kinase in human skeletal muscle. J Physiol 528: 221–226, 2000 [Abstract] [Google Scholar]
49. Wright DC, Geiger PC, Han DH, Jones TE, Holloszy JO. Calcium induces increases in peroxisome proliferator-activated receptor gamma coactivator-1alpha and mitochondrial biogenesis by a pathway leading to p38 mitogen-activated protein kinase activation. J Biol Chem 282: 18793–18799, 2007 [Abstract] [Google Scholar]
50. Wu H, Kanatous SB, Thurmond FA, Gallardo T, Isotani E, Bassel-Duby R, Williams RS. Regulation of mitochondrial biogenesis in skeletal muscle by CaMK. Science 296: 349–352, 2002 [Abstract] [Google Scholar]
51. Wu H, Rothermel B, Kanatous S, Rosenberg P, Naya FJ, Shelton JM, Hutcheson KA, DiMaio JM, Olson EN, Bassel-Duby R, Williams RS. Activation of MEF2 by muscle activity is mediated through a calcineurin-dependent pathway. EMBO J 20: 6414–6423, 2001 [Europe PMC free article] [Abstract] [Google Scholar]
52. Wu Z, Puigserver P, Andersson U, Zhang C, Adelmant G, Mootha V, Troy A, Cinti S, Lowell B, Scarpulla RC, Spiegelman BM. Mechanisms controlling mitochondrial biogenesis and respiration through the thermogenic coactivator PGC-1. Cell 98: 115–124, 1999 [Abstract] [Google Scholar]
53. Yan Z, Choi S, Liu X, Zhang M, Schageman JJ, Lee SY, Hart R, Lin L, Thurmond FA, Williams RS. Highly coordinated gene regulation in mouse skeletal muscle regeneration. J Biol Chem 278: 8826–8836, 2003 [Abstract] [Google Scholar]
54. Zong H, Ren JM, Young LH, Pypaert M, Mu J, Birnbaum MJ, Shulman GI. AMP kinase is required for mitochondrial biogenesis in skeletal muscle in response to chronic energy deprivation. Proc Natl Acad Sci USA 99: 15983–15987, 2002 [Europe PMC free article] [Abstract] [Google Scholar]

Articles from American Journal of Physiology - Cell Physiology are provided here courtesy of American Physiological Society

Citations & impact 


Impact metrics

Jump to Citations
Jump to Data

Citations of article over time

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.1152/ajpcell.00481.2009

Supporting
Mentioning
Contrasting
10
187
1

Article citations


Go to all (147) article citations

Data 


Data behind the article

This data has been text mined from the article, or deposited into data resources.

Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.

Funding 


Funders who supported this work.

NIAMS NIH HHS (1)