the co-activator pgc-1: from flies to mammals pgc-1 co...

24
THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co-activator family regulates muscle formation and mitochondrial biogenesis Mª Cruz Marín, Juan J. Arredondo & Rafael Garesse and Margarita Cervera* Instituto Investigaciones Biomédicas Alberto Sols. Facultad de Medicina, UAM-CSIC Running title: Drosophila co-activator PGC-1 *Address correspondence to: Margarita Cervera, Instituto Investigaciones Biomédicas Alberto Sols. Facultad de Medicina, UAM-CSIC. Arzobispo Morcillo 4, 28029 Madrid, SPAIN. Fax: +34 91 585 4401 Phone: +34 91 497 5402; email: [email protected]

Upload: others

Post on 29-May-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

THE CO-ACTIVATOR PGC-1: from flies to mammals

PGC-1 co-activator family regulates muscle formation and mitochondrial

biogenesis

Mª Cruz Marín, Juan J. Arredondo & Rafael Garesse and Margarita Cervera*

Instituto Investigaciones Biomédicas Alberto Sols. Facultad de Medicina, UAM-CSIC

Running title: Drosophila co-activator PGC-1

*Address correspondence to: Margarita Cervera, Instituto Investigaciones Biomédicas

Alberto Sols. Facultad de Medicina, UAM-CSIC. Arzobispo Morcillo 4, 28029 Madrid,

SPAIN. Fax: +34 91 585 4401 Phone: +34 91 497 5402; email:

[email protected]

Page 2: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

2

ABSTRACT

Transcriptional co-activators and co-repressors are emerging as physiological

integrators of energy metabolism and other biological processes. Tissues and organs

need to adapt to different critical physiological process as energetic metabolism,

apoptosis or ROS signaling. The correct number of structural and functionally

differentiated mitochondria mediates all these processes. In consequence, mitochondrial

biogenesis must be strictly regulated, a process that implies the coordinated regulation

of nuclear and mitochondrial genome gene expression. In mammals, PGC-1 co-

activators are essential to integrate the response of a network of specific transcription

factors involved in mitochondria biogenesis. Here, we focus on the role of the PGC-1

co-activator family. These proteins are key players in the control of organ-specific

biologic responses to the physiologic environment. Emphasis will be given to tissue-

specific regulatory features relevant to muscle. Our knowledge of the PGC-1 family in

vertebrates has augmented, but even now the molecular networks on which these co-

regulators are implicated, the PGC-1 response to the external signal and how molecular

pathways are functionally integrated to activate gene expression programs are mostly

unknown. In this sense, the studies of the function of the unique member of Drosophila

identified and the comparative analysis between flies and mammals will help in our

understanding of the integrative signals mediated by PGC-1 co-activator protein family.

Here we show that Drosophila PGC-1, Spargel/dPGC-1 may play also a crucial role in

mitochondrial biogenesis. Homozygous dPGC-1 null mutants are not fertile and present

strong muscle phenotype. These mutants are unable to fly and electron microscopy

studies of flight muscles from these animals show clear fiber degeneration. Very

interestingly, we have also found a strong reduction of both mitochondrial electron

density and size. Fly and mammals comparative analysis will be presented.

Page 3: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

3

INTRODUCTION

During development as animal grow, cells need to adapt to different physiological

process as energetic metabolism, apoptosis or ROS signaling. In these processes,

mitochondria have critical anabolic and catabolic functions in metabolizing nutrients

and in adapting their cellular physiology. The correct number of structural and

functionally differentiated mitochondria mediates all these actions. Therefore,

mitochondrial biogenesis must be strictly regulated and nuclear and mitochondrial

genome’s gene expression must be coordinated.

The last two decades have witnessed a tremendous expansion in our knowledge of the

mechanisms employed by eukaryotic cells to control gene activity. Transcriptional co-

activators and co-repressors are emerging as physiological integrators of energy

metabolism and other biological events [1, 2]. A common theme of these co-activator

regulatory networks it that, by co-activating multiple transcription factor targets they are

able to coordinate diverse biological processes that constitute biological responses [3,

4].

To learn about mitochondrial biogenesis, most studies focused on individual tissues that

show an enormous increase in mitochondrial activity and mass in response to external

stimuli, for example, during brown adipose tissue, muscle or perinatal heart maturation.

These studies led to identification and characterization of the PGC-1 family of

transcriptional co-activators that are potent inducers of mitochondrial biogenesis [5, 6].

A critical aspect of the PGC-1 co-activators is that they are highly versatile and have the

ability to interact with many different transcription factors. In fact, in vertebrates, the

PGC-1 proteins are inducible transcriptional co-activators orchestrating control of

cellular energy metabolism [4, 7]. Less is known about the role of the unique member of

the PGC-1 family present in the Drosophila model system, recently identify by our

group and others [8].

Although our knowledge of the PGC-1α in vertebrates has increased substantially, not

all the molecular pathways in the biological processes on which these co-regulators are

involved are known. Moreover, it is unclear how PGC-1α responds to the different

external stimuli and how molecular pathways are functionally integrated in different

gene expression programs.

Page 4: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

4

In flies, the homologue of PGC-1a, Spargel/dPGC-1, seems to play a critical role in

mitochondrial biogenesis in fat body and muscle, coordinating mitochondrial activity

and insulin signaling [8]. Drosophila is an excellent system to profound the knowledge

of mitochondria and muscle for three reasons: 1) Its powerful genetics combined with a

wealth of available cell and molecular biology techniques; 2) Genetic pathways that

guide basic biological processes, in Drosophila and in vertebrates, have remained

largely intact throughout evolution. This conservation enables rapid analysis of these

processes in vertebrates and 3) Habitually in flies, one single member represents a

functional gene family in vertebrates facilitating functional studies.

PGC-1 family members in vertebrates drive mitochondrial biogenesis through co-

activation of nuclear transcription factors. The majority of these factors are conserved in

flies. Thus, the transcription factors, Erect wing, homolog of NRF1, plays a critical role

in neural and muscle development; Delg, homolog of NRF2, is required during

oogenesis; NRGF, binds NRG sequences (nuclear regulatory gene elements) localized

in mitochondrial gene promoters and DREF is involved in DNA replication and cell

cycle control are conserved [8, 9]. At the moment, no Drosophila transcriptional co-

regulators have been characterized as physiological integrators of energy metabolism

and other biological processes. We think flies will help, in the near future, to unravel

some of the open and common questions about the exact role of PGC-1 in the

integration and coordination of the different regulatory networks in which this co-

activator is implicated.

This review will focus on the role of the PGC-1 co-activator family. These proteins are

key players in the control of organ-specific biologic responses to the physiologic milieu.

Emphasis will be given to tissue-specific regulatory features relevant to fat body and

muscle. Our knowledge of the PGC-1family in vertebrates has augmented, but even

now the molecular networks on which these co-regulators are implicated are largely

known. At the moment, the complete mechanisms of how PGC-1 responds to the

external signals and how molecular pathways are functionally integrated to activate

gene expression programs are unknown. In this sense, the studies of the function of the

unique member of Drosophila identified [8], and the comparative analysis between flies

and mammals will help in our understanding of the integrative signals mediated by

PGC-1 co-activator protein family. Fly and mammals comparative analysis will be

Page 5: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

5

presented.

RESULTS AND DISCUSSION

Molecular Conservation of PGC-1 co-activator family genes

In silico analysis has shown that PGC-1 family of co-activators is conserved in many

chordate species, including primates, rodents, ruminants, birds, amphibians and fishes

[5]. In vertebrates, three members compose this family; the first identified was PGC-

1α , characterized as a PPARγ -interacting protein and isolated from brown fat [10].

PGC-1α has two structural homologues, PGC-1β which shares an extensive sequence

homology with PGC-1α [11, 12], and PRC-related co-activator (PRC) with more

limited overall sequence similarity with PGC1α and PGC1β but sharing the key

structural features that define the family [13]. Detailed protein analysis reveals different

conserved functional domains (Figure 1). In the N-terminal acidic region, there is a

transcription activation domain that contains one nuclear hormone co-activator

signature motif (LXXLL) in PGC-1α and two of these motifs in PGC-1β that are crucial

for co-activation through certain nuclear receptors. In the central region of the PGC-1

proteins, a regulatory domain, mainly involved in co-activation through binding to

different mitochondrial transcriptional factors, exists. In the C-terminal region a

Ser/Arg-rich motif is situated close to a RNA binding domain [5].

In Drosophila melanogaster, a single homologue, annotated CG9809 and located in

chromosome 3R, has been identified [14]. This gene encodes a predicted protein of

1.088 amino acids and contains a RNA binding motif in the C-terminal region that

exhibits approximately 60% of homology with the same region in mammalian PGC-1 α

and β genes. Moreover, the Drosophila protein includes an acidic NH2-terminal

domain and an arginine-serine-rich domain common with its mammalian homologue

genes [14, 15] (Figure 1). Although the canonical LXXLL nuclear receptor binding

motif is not present, dPGC-1 contains a conserved variant of this motif FXXLL in the

C-terminal region which has been demonstrated to be involved in nuclear receptor

binding [16]. Thus, Drosophila and vertebrate PGC-1α proteins contain the functional

domain involved in nuclear receptor binding.

Page 6: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

6

FIGURE 1: Molecular conservation of mammalian and Drosophila PGC-1 co-activator family. In

the upper part the sequence homology of PRC, PGC-1β and PGC-1α proteins is shown. Color boxes

represent protein domains: activation domain (red), Arg/Ser rich domain (orange) and RNA binding

domain (blue). Black lines denote LXXL motifs (nuclear receptor binding sites). In the lower part, PGC-

1α and PGC-1 of Drosophila (dPGC-1) sequence homology is presented. Note the high percentage of

homology in RNA binding domain, RRM domain. No LXXL-like motifs have been identified in

Drosophila. Also note that sequence F(D/E)XLL is common to the carboxy-terminal domains of all PGC-

1 family members.

PGC1 proteins function as inducible transcriptional co-activators: molecular mechanism

The discovery of the first transcriptional co-activators and co-repressors for nuclear

hormone receptors and other transcription factors initiated a novel concept centered

around the control of specific genetic programs coordinated at the level of these proteins

[10]. Co-activators act interacting with various transcription factors, allowing for the

coordinated expression of gene sets in response to specific signals. Another important

and interesting implication of co-activators is the possibility to activate gene expression

in very specific contexts. The members of the inducible PGC-1 transcriptional co-

activator family posses a common function in mitochondrial physiology in addition to

control other specific programs. In the cells, PGC-1 gene expression is modified in

response to a variety of extracellular stimuli including temperature, energy deprivation

and availability of nutrients and grows factors [6, 17, 18]. However, these co-activators

exhibit differential specificities for transcription factors utilization, which in turn

modulates the physiological response orchestrated by each co-activator [7].

Page 7: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

7

The molecular mechanisms through which PGC-1α activates gene expression are

poorly understood. These co-activators form part of protein complexes whose concrete

protein identity and enzymatic activities are not completely known. Molecular studies,

mostly performed in cell cultures, indicate that PGC-1α and PGC-1β function through

recruiting chromatin-remodeling complexes that include those involved in histone

modification and nucleosome remodeling (reviewed in [18]). PGC-1α activates

transcription by physical interaction with p300 histone acetyltransferase domain

containing proteins as P300, CBP, SRC-1 and GCN5 as well as the mediator protein

complex, which is thought to recruit RNA polymerase II [19-22]. In the liver, where it

participates in the regulation of lipid metabolism, PGC-1α interacts with BRG1/BRM-

associated factor (BAF60a), a subunit of SW1/SNF nucleosome remodeling complex.

This interaction is required for the transcriptional function of PPARα (peroxisome

proliferator-activated receptor) and may be implicated in transcriptional function of

other nuclear hormone receptors [23, 24]. The recruitment of these histone acetyl-

transferase complexes containing PGC-1α to the proximity of the target-genes increases

histone acetylation, one of the first steps in transcriptional activation.

Another important aspect to understand is how these complexes may be spatially and

temporally assembled with PGC-1α to modulate and activate gene expression. Not

much is known about how these events occur, but modulation of transcription is

achieved through acetylation /deacetylation of PGC-1α itself. PGC-1 α has been found

in association with two additional different transcriptional complexes, the GCN5 and

TIP60 acetyl transferase complexes. The acetyl transferase GCN5 directly acetylates

PGC-1α at multiple lysine residues and the acetylated protein negatively regulates its

transcriptional activation function, at least in part through nuclear sub-localization [21].

The opposite role is played by lysine de-acetylation of PGC-1 by SIRT1 which

increases its function in response to cellular nutrients [25-28].

The model described above is perhaps simplistic and does not take into account other

additional PGC-1 post-translational modifications such as phosphorylation and

methylation whose precise functions are still unknown. Thus, PGC-1α phosphorylation

by the p38 MAPK and AMP kinase (AMPK) has been described [29-31]. Also,

methylation [32] and the novel O-linked β-N-acetylglucosamine (O-GlcNAc)

modifications on PGC-1α have been shown. The last, modulating the activity of FoxO

Page 8: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

8

transcription factors in liver culture cells [33]. On the other hand, the RNA-binding

domain of PGC-1α has been implicated in processing many mRNAs whose

transcription it initiates [34].

PCG-1α, PGC-1β and PRC act as transcriptional co-activators and function through

direct physical interaction with transcription factors directly bound to DNA promoter

regions. Thus, through their N-terminal region they interact with different hormone

nuclear receptors including PPARs, HNF4α, GR and ERRα while through their central

region bind transcription factors implicated in the activation of mitochondrial

biogenesis and respiratory function, such as NRF1, NRF2, TFAM, TBPs and MEF2C

[5, 7]. In addition, they interact with the RNA-processing C-terminal region with other

transcriptional factor group such as FoxO and YY1 involved in other pathways [35]. As

mentioned in the Introduction, the ability of PGC-1α to interact with different

transcription factors allows for the coordinated expression of gene sets in response to

specific signals (Figure 2).

In this context, in vertebrates, it has been described that NRF-1 transcription factor

regulates the expression of many genes involved in mitochondrial regulatory including

the majority of nuclear genes that encode subunits of the five respiratory complexes as

well as the expression of Tfam [36], and the two TFB genes [37], the main regulators of

mitochondrial transcription. Both NRF-1 and PGC-1α are up-regulated during adaptive

mitochondrial response of skeletal muscle to exercise training [38, 39] and in cultured

myotubes in response to increased calcium [40]. PGC-1β also is a potent co-activator of

NRF-1. An increase in PGC-1β expression in transgenic mice promotes, associated

with an increase of oxidative muscle fibers, a massive increase of mitochondrial

contend and high levels of mitochondrial respiratory genes transcription [41]. Thus,

PGC-1α , PGC-1β and PRC bind NRF-1 in order to trans-activate NRF-1 target genes.

Serum treatment of quiescent fibroblast induces PRC expression, followed by an

increase of TFAM, TFB1M and B2 genes, as well as nuclear and mitochondrial

encoded subunits of respiratory complexes. Functional binding sites for NRF-2 (the

human homolog of mouse GA- binding protein) have been localized to the promoters of

numerous genes related to respiratory chain, mainly COX subunits [42, 43], as well as

to TFAM [44],TFB1 and B2 promoters [37]. It has been speculated that coordinated

regulation of both NRF-1 and NRF-2 proteins mediated by PGC-1 family members

Page 9: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

9

could be a mechanism to ensure the correct expression of essential mitochondrial

proteins.

In the same context, the hormone nuclear receptor ERRα levels are high in oxidative

tissues and it modulates the activation of MCAD (medium-chain acyl-coenzyme A

deshydrogenase) promoter in the β -oxidation pathway [45]. ERRα has also been

implicated in PGC-1 induced mitochondrial biogenesis. Mootha and col. (2004),

through a computational approach to detect cis-regulatory motifs coupled with PGC-

1α -induced genome-wide transcriptional profiles, identified ERRα and GA-binding

protein α (NRF-2 α ) as key transcription factors regulating the OXPHOS pathway.

Cellular assays confirmed that both proteins were PGC-1α partners in muscle,

constituting a double-positive-feedback loop that drives the expression of many

OXPHOS genes [46].

FIGURE 2: Illustration summarizing PGC-1 functions in mammalian muscle cells. PGC-1 co-

activator integrates various physiological cues (red arrow) that are transduced into different signaling

cascades (yellow arrow) modulating the levels of PGC-1 mRNA and/or protein. In turn, PGC-1 mediated

Page 10: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

10

co-activation of several nuclear transcription factors orchestrates specific responses into the muscle cell

(green arrow).

In addition to its role in respiratory chain genes and mitochondrial transcription, PGC-

1α promotes mitochondrial oxidative functions by inducing the expression of genes of

the mitochondrial fatty acid oxidation and heme biosynthetic pathways. Ligand-

dependent binding PPARα to PGC-1α induce the trans-activation of PPARα target

genes [47]. Over-expression of PGC-1α induces MCAD gene expression through its

interaction with ERRα [45] and together with NRF-1 and FOXO1 (forkhead, box

family member) induces δ-aminolevulinate synthase, a critical enzyme in the heme

biosynthetic pathway [48]. Previously, it has been showed that control of glucose

oxidation through pyruvate deshydrogenase complex (PDC) could be regulated by

PGC-1α . PGC-1α over-expression induces, via ERRα , the transcriptional up-

regulation of PDK-4 (pyruvate deshydrogenase kinase), which inactivates PDC by

phosphorylation. In addition, PGC-1α over-expression effects were lost in MEFs

derived from a ERRα null mice [49]. Recent studies showed that PDK4 gene

expression is stimulated by thyroid hormone T(3) demonstrating that following T(3)

administration there is an increase in the association of PGC-1 α with the PDK4

promoter [50].

In flies, the majority of the transcription factors that interact with mammalian PCG-1α,

PGC1β and PRC have been identified. Thus, Drosophila transcription factor erect wing

(ewg) is the molecular homologue of the NRF-1 gene of vertebrates. ewg has been

reported as a critical transcription factor during muscle and nervous system

development. Surprisingly, to date, no ewg function related to mitochondrial biogenesis

or OXPHOS respiratory chain has been described [51]. Recently, Haussmann and col.

(2008) have showed that EWG neuronal expression restricts synaptic growth at

neuromuscular junctions. Using a functional genomics approach in late embryos they

demonstrated that EWG acts increasing the mRNA levels of genes involved in

transcriptional and post-transcriptional regulation of gene expression indicating the

existence of an extensive regulatory network. Then, EWG restricts synaptic growth by

integrating multiple cellular signaling pathways into an regulatory gene expression

network [52]. This result may be accounted with the role of NRF-1 as a critical

regulator in vertebrates. However, genes differentially regulated in ewg mutants were

not enriched in mitochondrial genes or genes involved in mitochondrial biogenesis.

Page 11: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

11

Also, mitochondrial content in synapses was not affected in these mutants. But

otherwise, in ewg mutants several genes that are differentially regulated are involved in

energy metabolism through the generation of ATP by inhibiting gluconeogenesis and

mobilizing lipids stores [52]. Preliminary studies carried out in our lab show that the

transcription of ewg gene is reduced in null flies. These results are in accordance to the

muscle phenotype observed in this null mutant (unpublished results). Since, PGC-1α

null mice have low levels of transcription of NMJ genes in nervous system [53], it is

worth to speculate that ewg role in NMJ transcriptional regulation in Drosophila could

be could be mediated by dPGC-1.

NRF2/GABPα in Drosophila, named Delg, is required in oogenesis and seems to have

an equivalent function to its mammalian homologue [54]. This factor is involved in the

transcriptional regulation of several mitochondrial genes. In addition, is required to

adjust mitochondrial mass depending on nutrients availability [9]. Previous studies

showed that dPGC-1 and Delg have a common role in regulating the expression of

many genes encoding mitochondrial proteins in mid-third instar larvae fat body.

Moreover dPGC-1 and Delg show synergistic effects over mitochondrial content as

demonstrated dPGC/Delg double mutant [8].

Additional Drosophila transcription factors as DREF [DRE (DNA replication-related

element)-binding factor] regulate the transcription a group of genes involved in

mitochondrial DNA replication and maintenance as well as several cell proliferation-

related genes [55]. NRGF (nuclear respiratory gene) is a transcription factor that

regulates the expression of genes with mitochondrial function. Among them, are found

subunits of the electron transport chain, proteins involved in the oxidative metabolism

and mitochondrial biogenesis. NRGF binds the NRG response elements which are

found in more than 50% of the promoters of these genes [56]. Currently, it has been

speculated that this factor could be putative co-regulators of dPGC-1 mediated

response. Future approaches will allow the identification of novel co-regulators

indispensable to determine molecular mechanism of dPGC-1 in Drosophila.

Physiological pathways implicated in external response to modulate PGC-1 expression

In response to external stimuli, physiological effectors pathways mediate changes in the

transcriptional expression of PGC-1 co-activators. The c-AMP-dependent pathway is

Page 12: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

12

one of several that direct the induction of PGC-1α in a number of tissues. c-AMP is

induced in response to thermal or fasting signals so that, c-AMP high levels lead to

activation of CREB family members, through its phosphorylation by protein kinase A

(PKA). The PGC-1α promoter has a potent c-AMP response element (CRE) that serves

as a target for CREB- mediated transcription activation. During energy deprivation

conditions (reduced ATP) such as starvation, ischemia and chronic metabolic stress,

AMP-activated protein kinase (AMPK) serves as an energy sensor for whole body

energy regulation [57]. AMPK is activated by decreases in both the ATP/AMP ratio and

phosphocreatine content. After activation it increases glucose transport, fatty acid

oxidation, and mitochondrial biogenesis. In skeletal muscle AMPK directly

phosphorylates PGC-1α and increases its expression through a feedback loop [31, 58].

Also, skeletal muscle adapts to chronic physical activity through the

calcium/calmodulin-dependent protein kinase IV (CaMKIV) and calcineurin A (CnA).

PGC-1α promoter is regulated by both CaMKIV and CnA activity, and activates PGC-

1α through the binding of c-AMP response element-binding. Furthermore, the calcium-

signaling pathway may result in a stable induction of PGC-1α , contributing to the

relatively stable nature of muscle fiber-type determination [59].

PGC-1α also is induced via cGMP-dependent signaling resulted from elevates levels of

nitric oxide (NO). Long-term exposure of cells in culture to low concentrations of NO

induces mitochondrial biogenesis. This process is mediated by cGMP, resulting from

NO-dependent activation of soluble guanylatecyclase (sGC), and involves increased

expression of PGC-1α , NRF-1 and Tfam [60]. In addition, studies in mice deficient in

endothelial NO synthase (eNOS) present reduced mitochondrial biogenesis in tissues

associated with reduced energy expenditure. NO/cGMP-dependent mitochondrial

biogenesis yields functionally active mitochondria, in terms of respiratory function and

metabolic activity [61].

Almost nothing is known in Drosophila about how dPGC-1 responds to the different

external stimuli and how molecular pathways are functionally integrated. Recently

microarray studies have shown that starvation, produces a subsequent reduction of

insulin pathway and a decrease of expression of genes encoding mitochondrial proteins

[14, 62]. The over-expression of insulin receptor (ISR-1) induces an increase of dPGC-1

transcription; additionally, transcriptional changes, produced in response to IRS-1

signaling, is attenuated in hypomorphicdPGC-1 mutant suggesting that dPGC-1 is a

Page 13: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

13

mediator of insulin signaling [8]. In other respect, which transcription factors are co-

activated by dPGC-1 to integrate different signal also remain unidentified.

Tissue protein levels and functions of PGC-1 co-activators

High levels of PGC-1 co-activators are detected in those tissues with high metabolic

activity including brown adipose tissue, liver, brain, heart and skeletal muscle. In

contrast with other transcriptional co-activators, PGC-1 proteins are highly induced in

response to developmental stage-specific and physiological cues [6, 17, 18]. PGC-1α

and PGC-1β are induced by cold exposure, exercise and fasting in tissue-specific

patterns [10, 39, 63, 64]. In addition, PGC-1α is induced in heart at birth and during

postnatal stages. These stages involve a great mitochondrial biogenesis response as well

as, in parallel, high energetic requirement [65-67]. In the case of Drosophila dPGC-1

the situation is comparable. The protein is expressed ubiquitously in flies but high levels

are detected in the same tissues than in vertebrates, such as skeletal and somatic

muscles, fat body, lymph gland and central nervous system

(http://insitu.fruitfly.org/cgi-bin/ex/insitu.pl). Curiously, high levels of mRNA are

detected also in oocytes, showing that dPGC-1 is supplied like maternal deposit [68].

Cells with different functions and energetic requirements have different mitochondrial

content, i.e. number of mitochondria, mtDNA content and/or structures. In addition,

cells may adapt mitochondrial function in response to cellular or environmental signals

such as hormones, grow factors, changes in temperature, physiological activity or

developmental cues. So, the control of mitochondrial biogenesis and function implies

complex and integrated regulatory networks that depend on the coordinated expression

of two genomes located in different sub-cellular compartments, the nucleus and the

mitochondria [69-71]. PGC-1 family members play a critical role in this inter-genomic

communication reviewed in [7].

First of all, in this situation, an important question to answer is if PGC1α co-activator is

indispensable for basal mitochondrial physiology. Studies of PGC-1α loss-of-function

in mice indicate that this co-activator is dispensable for the basal physiologic processes

of mitochondria metabolism or fetal development. However, these experiments also

demonstrated that mice have severely prejudiced their ability to respond to external

stimuli such as stress, cold exposure or starvation [72, 73]. In this context, therefore,

PGC-1a seems to be a molecular switch of the metabolic variation to external stimuli.

Page 14: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

14

PGC-1a not only plays an essential role in regulating energetic metabolism but also in

many other processes and events in which mitochondria plays a fundamental role.

The physiological significance of PGC-1α and PGC-1β in mitochondrial energy

metabolism has been studied by several approaches. Studies of over-expression in

mammalian culture cell or transgenic mouse, have demonstrated that both PGC-1α or

PGC-1β are able to activate, both, the expression of a number of genes important for

mitochondrial biogenesis, and respiratory function in adipocytes, cardiac myocytes and

myogenic cells [11, 30, 65, 74, 75]. These properties are shared by PGC-1β. PGC-1α

deficiency profoundly perturbs contractile function and leads heart failure [66]. In

human, the expression of PGC-1 and mitochondrial OXPHOS genes is significantly

reduced in skeletal muscle of diabetic patients, implicating a potential role for this

factor in the pathogenesis of insulin resistance [46]. Recently, studies in vivo and in

vitro demonstrated that PGC-1α stimulates a broad and powerful program of NMJ

(neuromuscular junctions)-linked gene expression in skeletal muscle [53]. They

concluded that PGC-1α to be a key integrator of neuromuscular activity in skeletal

muscle. Ectopic expression of PGC-1α in muscle results in increased mitochondrial

number and function as well as an increase in oxidative, fatigue-resistant muscle fibers.

PGC-1 co-activators mediate muscle external adaptive stimuli (Figure 2). Whole body

PGC-1α knock-out mice have a very complex phenotype but do not have a marked

skeletal muscle phenotype. They thus analyzed skeletal muscle-specific PGC-1αknock-

out mice to identify a specific role for PGC-1α in skeletal muscle function. These mice

exhibit a shift from oxidative type I and IIa toward type IIx and IIb muscle fibers.

Moreover, skeletal muscle-specific PGC-1αknock-out animals have reduced endurance

capacity and exhibit fiber damage and elevated markers of inflammation following

treadmill running. Their data demonstrate a critical role for PGC-1α in maintenance of

normal fiber type composition and of muscle fiber integrity following exertion. This

conversion also involves changes in the expression of specific-fiber genes such as

troponin I (slow) and myoglobin genes. Using culture cells, the same group has shown

that this regulation is mediated by PGC-1α co-activation of MEF2 which is a essential

muscle transcription factors involved in fiber type determination [30, 59].Similarly to

PGC-1α , transgenic expression of PGC-1β , causes a marked induction of IIX fibers,

which are oxidative but have "fast-twitch" biophysical properties. PGC-1β also co-

Page 15: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

15

activates MEF2 transcription factor to bind to type IIX myosin heavy chain (MHC)

promoter and activate its expression. PGC-1β transgenic muscle fibers are rich in

mitochondria and are highly oxidative, at least in part, due to co-activation by PGC-1 β

of ERRα and PPARα . Consequently, these transgenic animals can run for longer and at

higher workloads than wild-type animals [41]. Moreover, in vivo skeletal muscle

experiments have demonstrated that PGC-1α powerfully regulates VEGF (vascular-

endothelial-grow-factor) expression resulting in muscle angiogenesis augmentation.

PGC-1α null mice submit to ischemic insult failure to reconstitute blood flow in a

normal manner to the limb whereas transgenic expression of PGC-1α in skeletal muscle

is protective. Instead, PGC-1α co-activates ERR-α on conserved binding sites found in

the promoter VEGF gene. Thus, PGC-1α and ERR-α also control a novel angiogenic

pathway that delivers needed oxygen and substrates [76].

FIGURE 3: The absence of dPGC-1 leads to muscle fiber degeneration, mitochondrial morphology

and mitochondrial content in null dPGC-1 flies. Transmission electron microscopic analysis of thorax

sections from wt and null dPGC-1 flies are shown. A and B panels illustrate longitudinal sections and C

and D panels show transversal sections. A and C panels represent wild type flight muscle sections and B

Page 16: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

16

and D panels illustrate altered morphology and mitochondria alignment along the fibers in flight muscle

sections in dPGC -/- flies (arrows and asterisk). (Scale bars=1μm). M, mitochondria; Z, Z axis.

In the other hand, PGC-1β -deficient mice shows milder stress-induced phenotypes but

very similar to PGC-1α null mice [77, 78]. To complete this analysis, double deficient

PGC-1α and PGC-1β null mice were done and the analysis of the phenotype revealed

that they die shortly after birth as result of heart failure, including abnormal cardiac

maturation and block in mitochondrial biogenesis. These studies clearly demonstrate

that both proteins have overlapped targets and therefore, the absence of one of the

members may be partially compensated by the other factor but the absence of both

generates a strong phenotype.

Similar to PGC-1α deficiency mice, Drosophila dPGC-1 null flies are viable and

females are sterile. In addition, adult null flies showed locomotor alterations that

partially unable them to flight (unpublished results). The flight was reduced to 50%,

indicating an important malfunctioning of flight muscles. Muscle and mitochondria

morphologies in null flight muscles were examined to determine whether structural

abnormalities underlay the defects in muscle function. As shown in figure 3, the ultra-

structure of the flight muscles from wild type flies is extremely regular. The myofibrils

are circular in cross section and exhibit regular sarcomeric units that are aligned in

parallel between adjacent myofibrils in longitudinal section. Between these myofibrils a

highly packaged number of electron-dense mitochondria that exhibit the classical

double membrane structure with cristae visualized as invaginations of inner membrane

that extend into the matrix are observed together with a small number of nuclei. In

contrast, flight muscles of mutant null flies exhibit important abnormalities. Thus, most

of the muscles were severely disrupted and a high percentage of muscle fibers were

observed wavy, misalignment and degenerated. Myofibril diameters were variable and

mostly decreased and irregulars. An important increase of connective tissue was

visualized between muscle fibers, together with a modestly reduced number and size of

mitochondria relative to wt muscle control. In addition to being less abundant with

irregular shapes, these organelles exhibit slightly reduced electron density and, instead

of recognizable cristae, contain a granular internal structure. Swollen sarcoplasmic

reticulum-like structures were frequently observed. This muscle phenotype varies in

penetrance, some muscles were only moderately disrupted whereas other muscles in the

same fly are severely affected. Thus, the diminished expression of dPGC-1 protein

Page 17: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

17

generates clear muscle abnormalities. These include an increase of connective tissue,

changes in the mitochondria size and rearrangements in mitochondria and nuclei in

Drosophila flight muscles. To corroborate these data COX activity assays and NADH

assays were used to determine mitochondrial mass or OXPHOS activity. Related to fly

muscle changes, recently, Tiefenbock and others, [8] observed a diminution of adult

weight in a hypomorphic allele mutants. Additionally, genes involved in the formation,

maintaining and function of the fiber, including MEF2 and TnI genes, are also reduced

in these null mutants (unpublished results). These observations supplies locomotor

phenotype of dPGC-1 null mutant and also are in agreement with the role of PGC-1 co-

activators family in muscle adaptability function in vertebrates.

Regarding other functions of PGC-1α as modulator of other metabolic events, it also

activates gene regulatory programs involved in hepatic gluconeogenesis [79-81], muscle

glucose uptake [82, 83], adaptive thermogenesis and heme biosynthesis [48, 84]. Recent

studies show that PGC-1α and PGC-1β control mitochondrial capacity in an additive

and independent manner in neurons. These evidences demonstrate that activation or

over-expression of members of the PGC-1 family could be used to compensate for

neuronal mitochondrial loss [85].

The role of the third member of the family, PRC (PGC-1 related co-activator), has been

investigated by stably silencing PRC expression with two different shRNAs. Complete

PRC silencing resulted in a severe inhibition of respiratory energy production associated

with the reduced expression and assembly of respiratory complex. In addition to the

respiratory defect, PRC silencing was also implicated in the inhibition of cell cycle

progression. Global changes in gene expression were consistent with both mitochondrial

and growth phenotypes[86-88]. Recent studies of PRC-loss- function support a role for

PRC in the integration of pathways directing mitochondrial respiratory function and cell

growth [89].

CONCLUDING REMARKS

In this review we have exposed the important role of the PGC-1 co-activator family as

metabolic sensors. These proteins are central players in the control of organ-specific

responses to a wide variety of physiological stimuli as well as different kinds of stress.

We have shown that PGC-1α is, both in mammals and flies, one of the key regulatory

Page 18: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

18

factors in tissues with a high energetic demand. Phenotypic adaptations to different

signaling pathways, including physical activity, metabolic demands, neuronal

innervations or ischemia converge on PGC-1 α regulation through its interaction with

diverse transcription factors. These interactions increase mitochondrial biogenesis and

function leading to a higher ATP production through both, OXPHOS metabolism and

fatty acid β -oxidation. PGC-1α has a particularly relevant role in muscle development

and adaptation where this co-factor regulates the expression of fiber specific and NMJ

post-synaptic genes. On the other hand, in response to ischemia, PGC-1α regulates

angiogenesis in muscle, accelerating the recovery of blood flow. In consequence, the

therapeutic potential of PGC-1α against diseases associated with deregulated muscle

function is broad. In fact, PGC-1α has shown an anti-atrophic effect in different

experimental systems.

ACKNOWLEDGMENTS

This research was supported by grants BFU2007-61711BMC from the DGICYT

(Spanish Ministry of Education, Culture and Sports) and S2006-GEN0269, MITOLAB

network, from the Comunidad de Madrid, Spain. We thank Dr Lucia Guerrero for

critical reading of the manuscript and Dr Mark Sefton from BiomedRed for checking

the English usage in this manuscript. Dr. Mª Cruz Marin, postdoctoral researcher, is

supported by MITOLAB network from the Comunidad de Madrid, Spain.

Page 19: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

19

REFERENCES

1. Christian, M., White, R. and Parker, M.G., 2006, Trends Endocrinol Metab, 17,

243.

2. Feige, J.N. and Auwerx, J., 2007, Trends Cell Biol, 17, 292.

3. Handschin, C. and Spiegelman, B.M., 2008, Nature, 454, 463.

4. Spiegelman, B.M. and Heinrich, R., 2004, Cell, 119, 157.

5. Lin, J., Handschin, C. and Spiegelman, B.M., 2005, Cell Metab, 1, 361.

6. Handschin, C. and Spiegelman, B.M., 2006, Endocr Rev, 27, 728.

7. Scarpulla, R.C., 2008, Physiol Rev, 88, 611.

8. Tiefenbock, S.K., Baltzer, C., Egli, N.A. and Frei, C., Embo J, 29, 171.

9. Baltzer, C., Tiefenbock, S.K., Marti, M. and Frei, C., 2009, PLoS One, 4, e6935.

10. Puigserver, P., Wu, Z., Park, C.W., Graves, R., Wright, M., and Spiegelman,

B.M., 1998, Cell, 92, 829.

11. Lin, J., Puigserver, P., Donovan, J., Tarr, P., and Spiegelman, B.M., 2002, J Biol

Chem, 277, 1645.

12. Kressler, D., Schreiber, S.N., Knutti, D. and Kralli, A., 2002, J Biol Chem, 277,

13918.

13. Andersson, U. and Scarpulla, R.C., 2001, Mol Cell Biol, 21, 3738.

14. Gershman, B., Puig, O., Hang, L., Peitzsch, R.M., Tatar, M., and Garofalo, R.S.,

2007, Physiol Genomics, 29, 24.

15. Lasko, P., 2000, J Cell Biol, 150, F51.

16. Huang, N., vom Baur, E., Garnier, J.M., Lerouge, T., Vonesch, J.L., Lutz, Y.,

Chambon, P., and Losson, R., 1998, Embo J, 17, 3398.

17. Finck, B.N. and Kelly, D.P., 2006, J Clin Invest, 116, 615.

18. Lin, J.D., 2009, Mol Endocrinol, 23, 2.

19. Puigserver, P., Adelmant, G., Wu, Z., Fan, M., Xu, J., O'Malley, B., and

Spiegelman, B.M., 1999, Science, 286, 1368.

20. Wallberg, A.E., Yamamura, S., Malik, S., Spiegelman, B.M., and Roeder, R.G.,

2003, Mol Cell, 12, 1137.

21. Lerin, C., Rodgers, J.T., Kalume, D.E., Kim, S.H., Pandey, A., and Puigserver,

P., 2006, Cell Metab, 3, 429.

Page 20: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

20

22. Kelly, T.J., Lerin, C., Haas, W., Gygi, S.P., and Puigserver, P., 2009, J Biol

Chem, 284, 19945.

23. Li, S., Liu, C., Li, N., Hao, T., Han, T., Hill, D.E., Vidal, M., and Lin, J.D.,

2008, Cell Metab, 8, 105.

24. Debril, M.B., Gelman, L., Fayard, E., Annicotte, J.S., Rocchi, S., and Auwerx,

J., 2004, J Biol Chem, 279, 16677.

25. Baur, J.A., Pearson, K.J., Price, N.L., Jamieson, H.A., Lerin, C., Kalra, A.,

Prabhu, V.V., Allard, J.S., Lopez-Lluch, G., Lewis, K., Pistell, P.J., Poosala, S.,

Becker, K.G., Boss, O., Gwinn, D., Wang, M., Ramaswamy, S., Fishbein, K.W.,

Spencer, R.G., Lakatta, E.G., Le Couteur, D., Shaw, R.J., Navas, P., Puigserver,

P., Ingram, D.K., de Cabo, R., and Sinclair, D.A., 2006, Nature, 444, 337.

26. Lagouge, M., Argmann, C., Gerhart-Hines, Z., Meziane, H., Lerin, C., Daussin,

F., Messadeq, N., Milne, J., Lambert, P., Elliott, P., Geny, B., Laakso, M.,

Puigserver, P., and Auwerx, J., 2006, Cell, 127, 1109.

27. Gerhart-Hines, Z., Rodgers, J.T., Bare, O., Lerin, C., Kim, S.H., Mostoslavsky,

R., Alt, F.W., Wu, Z., and Puigserver, P., 2007, Embo J, 26, 1913.

28. Rodgers, J.T., Lerin, C., Gerhart-Hines, Z. and Puigserver, P., 2008, FEBS Lett,

582, 46.

29. Fan, M., Rhee, J., St-Pierre, J., Handschin, C., Puigserver, P., Lin, J., Jaeger, S.,

Erdjument-Bromage, H., Tempst, P., and Spiegelman, B.M., 2004, Genes Dev,

18, 278.

30. Lin, J., Wu, H., Tarr, P.T., Zhang, C.Y., Wu, Z., Boss, O., Michael, L.F.,

Puigserver, P., Isotani, E., Olson, E.N., Lowell, B.B., Bassel-Duby, R., and

Spiegelman, B.M., 2002, Nature, 418, 797.

31. Jager, S., Handschin, C., St-Pierre, J. and Spiegelman, B.M., 2007, Proc Natl

Acad Sci U S A, 104, 12017.

32. Teyssier, C., Ma, H., Emter, R., Kralli, A., and Stallcup, M.R., 2005, Genes

Dev, 19, 1466.

33. Housley, M.P., Udeshi, N.D., Rodgers, J.T., Shabanowitz, J., Puigserver, P.,

Hunt, D.F., and Hart, G.W., 2009, J Biol Chem, 284, 5148.

34. Monsalve, M., Wu, Z., Adelmant, G., Puigserver, P., Fan, M., and Spiegelman,

B.M., 2000, Mol Cell, 6, 307.

35. Giguere, V., 2008, Endocr Rev, 29, 677.

36. Virbasius, J.V. and Scarpulla, R.C., 1994, Proc Natl Acad Sci U S A, 91, 1309.

Page 21: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

21

37. Gleyzer, N., Vercauteren, K. and Scarpulla, R.C., 2005, Mol Cell Biol, 25, 1354.

38. Baar, K., Wende, A.R., Jones, T.E., Marison, M., Nolte, L.A., Chen, M., Kelly,

D.P., and Holloszy, J.O., 2002, Faseb J, 16, 1879.

39. Pilegaard, H., Saltin, B. and Neufer, P.D., 2003, J Physiol, 546, 851.

40. Ojuka, E.O., Jones, T.E., Han, D.H., Chen, M., and Holloszy, J.O., 2003, Faseb

J, 17, 675.

41. Arany, Z., Lebrasseur, N., Morris, C., Smith, E., Yang, W., Ma, Y., Chin, S.,

and Spiegelman, B.M., 2007, Cell Metab, 5, 35.

42. Virbasius, J.V., Virbasius, C.A. and Scarpulla, R.C., 1993, Genes Dev, 7, 380.

43. Ongwijitwat, S. and Wong-Riley, M.T., 2005, Gene, 360, 65.

44. Larsson, N.G., Wang, J., Wilhelmsson, H., Oldfors, A., Rustin, P., Lewandoski,

M., Barsh, G.S., and Clayton, D.A., 1998, Nat Genet, 18, 231.

45. Huss, J.M., Kopp, R.P. and Kelly, D.P., 2002, J Biol Chem, 277, 40265.

46. Mootha, V.K., Handschin, C., Arlow, D., Xie, X., St Pierre, J., Sihag, S., Yang,

W., Altshuler, D., Puigserver, P., Patterson, N., Willy, P.J., Schulman, I.G.,

Heyman, R.A., Lander, E.S., and Spiegelman, B.M., 2004, Proc Natl Acad Sci

U S A, 101, 6570.

47. Vega, R.B., Huss, J.M. and Kelly, D.P., 2000, Mol Cell Biol, 20, 1868.

48. Handschin, C., Lin, J., Rhee, J., Peyer, A.K., Chin, S., Wu, P.H., Meyer, U.A.,

and Spiegelman, B.M., 2005, Cell, 122, 505.

49. Wende, A.R., Huss, J.M., Schaeffer, P.J., Giguere, V., and Kelly, D.P., 2005,

Mol Cell Biol, 25, 10684.

50. Attia, R.R., Connnaughton, S., Boone, L.R., Wang, F., Elam, M.B., Ness, G.C.,

Cook, G.A., and Park, E.A., J Biol Chem, 285, 2375.

51. DeSimone, S.M. and White, K., 1993, Mol Cell Biol, 13, 3641.

52. Haussmann, I.U., White, K. and Soller, M., 2008, Genome Biol, 9, R73.

53. Handschin, C., Kobayashi, Y.M., Chin, S., Seale, P., Campbell, K.P., and

Spiegelman, B.M., 2007, Genes Dev, 21, 770.

54. Hsu, T., 2000, Oncogene 19, 6409.

55. Fernández-Moreno, M.A., Bruni, F., Adán, C., Hernández-Sierra, R., Polosa,

P.L., Cantatore, P., Garesse, R., and Roberti, M., 2009, Biochem J., 418,

56. Echevarría, L., Sánchez-Martínez, A., Clemente, P., Hernández-Sierra, R.,

Fernández-Moreno, M.A., and Garesse, R., 2008, Nuclear-mitochondrial

Crosstalk: Dual Genome Control of Mitochondrial Biogenesis and Metabolism.

Page 22: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

22

Landes Bioscience and Springer Science+Business Media (ed), G. Pesole, D.

Horner & M. Sardiello, 9.

57. Zong, H., Ren, J.M., Young, L.H., Pypaert, M., Mu, J., Birnbaum, M.J., and

Shulman, G.I., 2002, Proc Natl Acad Sci U S A, 99, 15983.

58. Akimoto, T., Pohnert, S.C., Li, P., Zhang, M., Gumbs, C., Rosenberg, P.B.,

Williams, R.S., and Yan, Z., 2005, J Biol Chem, 280, 19587.

59. Handschin, C., Rhee, J., Lin, J., Tarr, P.T., and Spiegelman, B.M., 2003, Proc

Natl Acad Sci U S A, 100, 7111.

60. Nisoli, E., Clementi, E., Paolucci, C., Cozzi, V., Tonello, C., Sciorati, C.,

Bracale, R., Valerio, A., Francolini, M., Moncada, S., and Carruba, M.O., 2003,

Science, 299, 896.

61. Nisoli, E., Falcone, S., Tonello, C., Cozzi, V., Palomba, L., Fiorani, M.,

Pisconti, A., Brunelli, S., Cardile, A., Francolini, M., Cantoni, O., Carruba,

M.O., Moncada, S., and Clementi, E., 2004, Proc Natl Acad Sci U S A, 101,

16507.

62. Teleman, A.A., Hietakangas, V., Sayadian, A.C. and Cohen, S.M., 2008, Cell

Metab, 7, 21.

63. Benton, C.R., Wright, D.C. and Bonen, A., 2008, Appl Physiol Nutr Metab, 33,

843.

64. Mathai, A.S., Bonen, A., Benton, C.R., Robinson, D.L., and Graham, T.E.,

2008, J Appl Physiol, 105, 1098.

65. Lehman, J.J., Barger, P.M., Kovacs, A., Saffitz, J.E., Medeiros, D.M., and

Kelly, D.P., 2000, J Clin Invest, 106, 847.

66. Arany, Z., He, H., Lin, J., Hoyer, K., Handschin, C., Toka, O., Ahmad, F.,

Matsui, T., Chin, S., Wu, P.H., Rybkin, II, Shelton, J.M., Manieri, M., Cinti, S.,

Schoen, F.J., Bassel-Duby, R., Rosenzweig, A., Ingwall, J.S., and Spiegelman,

B.M., 2005, Cell Metab, 1, 259.

67. Finck, B.N. and Kelly, D.P., 2007, Circulation, 115, 2540.

68. Chintapalli, V.R., Wang, J. and Dow, J.A., 2007, Nat Genet, 39, 715.

69. Cuezva, J.M., Ostronoff, L.K., Ricart, J., Lopez de Heredia, M., Di Liegro,

C.M., and Izquierdo, J.M., 1997, J Bioenerg Biomembr, 29, 365.

70. Enriquez, J.A., Fernandez-Silva, P. and Montoya, J., 1999, Biol Chem, 380, 737.

71. Garesse, R. and Vallejo, C.G., 2001, Gene, 263, 1.

Page 23: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

23

72. Lin, J., Wu, P.H., Tarr, P.T., Lindenberg, K.S., St-Pierre, J., Zhang, C.Y.,

Mootha, V.K., Jager, S., Vianna, C.R., Reznick, R.M., Cui, L., Manieri, M.,

Donovan, M.X., Wu, Z., Cooper, M.P., Fan, M.C., Rohas, L.M., Zavacki, A.M.,

Cinti, S., Shulman, G.I., Lowell, B.B., Krainc, D., and Spiegelman, B.M., 2004,

Cell, 119, 121.

73. Leone, T.C., Lehman, J.J., Finck, B.N., Schaeffer, P.J., Wende, A.R., Boudina,

S., Courtois, M., Wozniak, D.F., Sambandam, N., Bernal-Mizrachi, C., Chen,

Z., Holloszy, J.O., Medeiros, D.M., Schmidt, R.E., Saffitz, J.E., Abel, E.D.,

Semenkovich, C.F., and Kelly, D.P., 2005, PLoS Biol, 3, e101.

74. Wu, Z., Puigserver, P., Andersson, U., Zhang, C., Adelmant, G., Mootha, V.,

Troy, A., Cinti, S., Lowell, B., Scarpulla, R.C., and Spiegelman, B.M., 1999,

Cell, 98, 115.

75. St-Pierre, J., Lin, J., Krauss, S., Tarr, P.T., Yang, R., Newgard, C.B., and

Spiegelman, B.M., 2003, J Biol Chem, 278, 26597.

76. Arany, Z., Foo, S.Y., Ma, Y., Ruas, J.L., Bommi-Reddy, A., Girnun, G.,

Cooper, M., Laznik, D., Chinsomboon, J., Rangwala, S.M., Baek, K.H.,

Rosenzweig, A., and Spiegelman, B.M., 2008, Nature, 451, 1008.

77. Lelliott, C.J., Medina-Gomez, G., Petrovic, N., Kis, A., Feldmann, H.M.,

Bjursell, M., Parker, N., Curtis, K., Campbell, M., Hu, P., Zhang, D., Litwin,

S.E., Zaha, V.G., Fountain, K.T., Boudina, S., Jimenez-Linan, M., Blount, M.,

Lopez, M., Meirhaeghe, A., Bohlooly, Y.M., Storlien, L., Stromstedt, M.,

Snaith, M., Oresic, M., Abel, E.D., Cannon, B., and Vidal-Puig, A., 2006, PLoS

Biol, 4, e369.

78. Sonoda, J., Mehl, I.R., Chong, L.W., Nofsinger, R.R., and Evans, R.M., 2007,

Proc Natl Acad Sci U S A, 104, 5223.

79. Puigserver, P., Rhee, J., Donovan, J., Walkey, C.J., Yoon, J.C., Oriente, F.,

Kitamura, Y., Altomonte, J., Dong, H., Accili, D., and Spiegelman, B.M., 2003,

Nature, 423, 550.

80. Koo, S.H., Satoh, H., Herzig, S., Lee, C.H., Hedrick, S., Kulkarni, R., Evans,

R.M., Olefsky, J., and Montminy, M., 2004, Nat Med, 10, 530.

81. Zhu, L.L., Liu, Y., Cui, A.F., Shao, D., Liang, J.C., Liu, X.J., Chen, Y., Gupta,

N., Fang, F.D., and Chang, Y.S., Am J Physiol Endocrinol Metab,

Page 24: THE CO-ACTIVATOR PGC-1: from flies to mammals PGC-1 co ...digital.csic.es/bitstream/10261/78180/4/Thecoactivator PGC1.pdf · FIGURE 1: Molecular conservation of mammalian and PGCDrosophila-1

24

82. Wende, A.R., Schaeffer, P.J., Parker, G.J., Zechner, C., Han, D.H., Chen, M.M.,

Hancock, C.R., Lehman, J.J., Huss, J.M., McClain, D.A., Holloszy, J.O., and

Kelly, D.P., 2007, J Biol Chem, 282, 36642.

83. Zhu, L., Sun, G., Zhang, H., Zhang, Y., Chen, X., Jiang, X., Jiang, X., Krauss,

S., Zhang, J., Xiang, Y., and Zhang, C.Y., 2009, PLoS One, 4, e4182.

84. Handschin, C., Chin, S., Li, P., Liu, F., Maratos-Flier, E., Lebrasseur, N.K.,

Yan, Z., and Spiegelman, B.M., 2007, J Biol Chem, 282, 30014.

85. Wareski, P., Vaarmann, A., Choubey, V., Safiulina, D., Liiv, J., Kuum, M., and

Kaasik, A., 2009, J Biol Chem, 284, 21379.

86. Scarpulla, R.C., 2008, Ann N Y Acad Sci, 1147, 321.

87. Vercauteren, K., Pasko, R.A., Gleyzer, N., Marino, V.M., and Scarpulla, R.C.,

2006, Mol Cell Biol, 26, 7409.

88. Vercauteren, K., Gleyzer, N. and Scarpulla, R.C., 2008, J Biol Chem, 283,

12102.

89. Vercauteren, K., Gleyzer, N. and Scarpulla, R.C., 2009, J Biol Chem, 284, 2307.