REVIEW
Transcriptional regulatory circuits
controlling mitochondrial biogenesis
and function
Daniel P. Kelly
1,3
and Richard C. Scarpulla
2
1
Center for Cardiovascular Research, Departments of Medicine, Molecular Biology & Pharmacology, and Pediatrics,
Washington University School of Medicine, St. Louis, Missouri 63119, USA;
2
Department of Cell and Molecular Biology,
Northwestern Medical School, Chicago, Illinois 60611, USA
We are witnessing a period of renewed interest in the
biology of the mitochondrion. The mitochondrion serves
a critical function in the maintenance of cellular energy
stores, thermogenesis, and apoptosis. Moreover, alter-
ations in mitochondrial function contribute to several
inherited and acquired human diseases and the aging
process. This review summarizes our understanding of
the transcriptional regulatory mechanisms involved in
the biogenesis and energy metabolic function of mito-
chondria in higher organisms.
The mitochondrial genome
A defining feature of eukaryotic cells is that they contain
nuclear and mitochondrial genomes sequestered into
distinct subcellular compartments. The mitochondrial
genetic system is comprised of a circular DNA genome
(mtDNA, 16.5 kb in vertebrates; Fig. 1), the enzymes
required for its transcription and replication, and the pro-
tein synthetic machinery necessary for the translation of
13 mitochondrial mRNAs (for review, see Garesse and
Vallejo 2001). These mRNAs, which account for the en-
tire protein-coding capacity of mtDNA, encode essential
subunits of respiratory complexes I, III, IV, and V. The
extrusion of protons through complexes I, III, and IV is
coupled to the sequential transfer of electrons to a series
of carriers of increasing redox potential resulting in an
electrochemical proton gradient across the inner mem-
brane. Complex V, comprised of an ATPase coupled to
an inner membrane proton channel, can dissipate the
proton gradient in the synthesis of ATP or can couple
proton pumping to ATP hydrolysis to maintain the gra-
dient. mtDNA also encodes for two ribosomal and 22
transfer RNAs, required for translation by mitoribo-
somes within the matrix.
The limited coding capacity of mtDNA necessitates
that nuclear genes make a major contribution to mito-
chondrial metabolic systems and molecular architecture
(Garesse and Vallejo 2001). One major class of nuclear
genes contributes catalytic and auxiliary proteins to the
mitochondrial enzyme systems. For example, the major-
ity of the 100 or so subunits of the respiratory apparatus
are nucleus-encoded. In addition, nucleus-encoded meta-
bolic enzymes necessary for the oxidation of pyruvate,
fatty acids (-oxidation cycle), and acetyl-CoA (tricar-
boxylic acid cycle), the biosynthesis of certain amino
acids, and the manufacture of heme, among others,
are localized to the mitochondrion. A second class of
nuclear genes encodes protein import and assembly fac-
tors. A third class contributes key proteins that are re-
quired for the replication and expression of the mito-
chondrial genome including nucleic acid polymerases,
RNA processing enzymes, transcription and replication
factors as well as tRNA-synthetases, translation factors,
and ribosomal subunits. Thus, the program regulating
mitochondrial biogenesis involves the coordinate ac-
tions of nuclear and mitochondrial genes.
Regulatory proteins involved in mitochondrial
gene transcription
In yeast, mtDNA transcription is initiated at 20 tran-
scriptional units throughout the genome (for review, see
Poyton and McEwen 1996). In vertebrates, transcription
is initiated bidirectionally at two promoters, P
H
and P
L
for heavy (H) and light strands (L), respectively, within
the D-loop regulatory region (Shadel and Clayton 1997;
Clayton 2000). The D-loop is the longest noncoding re-
gion in vertebrate mtDNA and contains, in addition to
P
H
and P
L
, the H-strand replication origin (O
H
; Fig. 1). In
the “strand asymmetric model” of mtDNA replication,
the RNA transcript initiated at P
L
is cleaved in the vi-
cinity of three evolutionarily conserved sequence blocks
(CSB I, II, and III), and H-strand replication is initiated at
the sites of these cleavages (Bogenhagen and Clayton
2003). Thus, transcription is coupled to DNA replication
3
Corresponding author.
E-MAIL [email protected]; FAX (314) 362-0186.
Article and publication are at http://www.genesdev.org/cgi/doi/10.1101/
gad.1177604.
GENES & DEVELOPMENT 18:357–368 © 2004 by Cold Spring Harbor Laboratory Press ISSN 0890-9369/04; www.genesdev.org 357
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
and the sites of RNA cleavage are transition sites be-
tween RNA and DNA synthesis. A decision must be
made to continue transcription through the CSBs or to
truncate the nascent RNA to initiate DNA replication.
After DNA synthesis begins, the nascent strand is often
terminated downstream from a conserved element re-
ferred to as a termination-associated sequence (TAS).
This event accounts for the triple-stranded D-loop struc-
ture and may be important in controlling mtDNA levels
(Brown and Clayton 2002).
The mitochondrial H- and L-strand transcriptional
units differ from most nuclear genes in that they are
polygenic. In addition to the RNA primer for H-strand
replication, P
L
also directs the synthesis of a transcript
that is processed to one mRNA and eight of the 22 tRNAs.
The polygenic transcript directed by P
H
is processed to
14 tRNAs, 12 mRNAs, and the two rRNAs (for review,
see Garesse and Vallejo 2001). Both promoters share a
critical upstream enhancer that serves as the recognition
site for mitochondrial transcription factor A or Tfam
(previously mtTF-1 and mtTFA), an HMG-box protein
that stimulates transcription through specific binding to
upstream enhancers. Like other HMG proteins, Tfam
can bend and unwind DNA, properties potentially linked
to its ability to stimulate transcription upon binding
DNA (Fisher et al. 1992). In addition to specific promoter
recognition, Tfam binds nonspecific DNA with high af-
finity. This property along with its abundance in mito-
chondria suggests that it plays a role in the stabilization
and maintenance of the mitochondrial chromosome
through its phased binding to nonpromoter sites (Parisi
et al. 1993).
Several lines of evidence indicate that Tfam is required
for mtDNA replication and maintenance. Tfam knock-
out mice display embryonic lethality and depletion of
mtDNA (Larsson et al. 1998). In addition, Tfam levels
correlate well with increased mtDNA in ragged-red
muscle fibers and decreased mtDNA levels in mtDNA-
depleted cells (Larsson et al. 1994; Poulton et al. 1994).
ABF2, a related HMG-box yeast factor, is required for
mtDNA maintenance and respiratory competence (Diff-
ley and Stillman 1991). Expression of Tfam can comple-
ment an ABF2 deficiency in yeast, suggesting that the
two proteins are functionally homologous (Parisi et al.
1993). Interestingly, despite this functional complemen-
tation, ABF2 lacks an activation domain present in Tfam
and does not stimulate transcription (Dairaghi et al.
1995).
Significant progress has been made in the character-
ization of the mtDNA transcription initiation machin-
ery. A vertebrate mitochondrial RNA polymerase and a
specificity factor that are required for mitochondrial-spe-
cific initiation were initially identified and characterized
in Xenopus laevis (Antoshechkin and Bogenhagen 1995;
Bogenhagen 1996). Although purification of the human
polymerase has been elusive, a human cDNA that en-
codes a protein with sequence similarity to yeast mito-
chondrial and phage polymerases has been identified in
database screenings (Tiranti et al. 1997). The encoded
protein localizes to mitochondria, suggesting that it is a
bona fide mitochondrial polymerase. A human mito-
chondrial transcription factor B (h-mtTFB) cDNA has
also been isolated, and the encoded protein has proper-
ties consistent with it being a functional homolog of the
Figure 1. Human mitochondrial DNA (mtDNA). The
genomic organization and structural features of human
mtDNA are depicted in a circular genomic map. The
D-loop regulatory region is expanded and shown above.
Protein coding and rRNA genes are interspersed with
22 tRNA genes (denoted by the single-letter amino acid
code). The D-loop regulatory region contains the L- and
H-strand promoters (P
L
and P
H
, respectively) along with
the origin of H-strand replication (O
H
). mtDNA trans-
cription complexes containing mitochondrial RNA poly-
merase, Tfam, and TFB are depicted in the expanded
D-loop along with the conserved sequence blocks (CSB
I, II, and III). The origin of L-strand replication (O
L
)is
displaced by approximately two-thirds of the genome
within a cluster of five tRNA genes. Protein-coding
genes include cytochrome oxidase (COX) subunits 1, 2,
and 3; NADH dehydrogenase (ND) subunits 1, 2, 3, 4,
4L, 5, and 6; ATP synthase (ATPS) subunits 6 and 8; and
cytochrome b (Cytb). ND6 and the eight tRNA genes
encoded on the L-strand are in bold type and under-
lined; all other genes are encoded on the H-strand.
Kelly and Scarpulla
358 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
yeast specificity factor, sc-mtTFB (McCulloch et al.
2002). The protein is localized to mitochondria and can
bind DNA and stimulate transcription from an L-strand
promoter in vitro. Subsequently, two isoforms of h-mtTFB,
termed TFB1 and 2, were identified (Falkenberg et al.
2002). TFB1 is identical to the initial h-mtTFB isolate.
Like the yeast factor, both TFBs share sequence similari-
ties with rRNA dimethyltransferases, although the simi-
larity between TFB2 and this class of enzymes is weaker
than that of TFB1. Both TFB isoforms can support spe-
cific initiation from mitochondrial promoters in an in
vitro system containing purified recombinant proteins.
In this system, the TFB-dependent activation of tran-
scription depends on mitochondrial RNA polymerase
and Tfam (Fig. 1). Both TFBs interact with mitochondrial
RNA polymerase, but TFB1 has about one-tenth the
transcriptional activity of TFB2. In addition to binding
mitochondrial RNA polymerase, TFB1 also contacts the
C-terminal domain of Tfam (McCulloch and Shadel
2003). The region of contact between TFB1 and Tfam is
essential for transcriptional activation and corresponds
to a 29-amino-acid domain that was previously identi-
fied as a Tfam activation domain (Dairaghi et al. 1995).
This reinforces the distinction between Tfam and the
yeast HMG-box protein ABF2, which, like Tfam, is re-
quired for mtDNA maintenance but does not function as
a transcription factor.
Transcriptional regulators of nuclear encoded
mitochondrial proteins: the critical role of nuclear
respiratory factors 1 and 2
The cytochrome c and cytochrome oxidase genes have
served as the prototypes for identifying regulatory factors
that act on nuclear respiratory genes from both yeast and
mammalian cells. Early work in yeast demonstrated that
transcriptional regulation of the major cytochrome c iso-
form, CYC1, was mediated by oxygen and carbon sources
through the upstream activation sites, UAS1 and UAS2.
This work has been the subject of excellent reviews to
which the reader is referred for original citations (Zi-
tomer and Lowry 1992; Poyton and McEwen 1996).
The identification of nucleus-encoded transcription
factors required for the expression of the respiratory ap-
paratus in mammalian cells also began with the charac-
terization of the cytochrome c gene (for reviews, see
Scarpulla 1997, 1999). Interestingly, the mammalian cy-
tochrome c promoter has multiple recognition sites for
transcription factors that bear no obvious relationship to
those identified in yeast (Evans and Scarpulla 1988). A
potent cis-acting element, localized to the first intron,
consists of tandem Sp1 recognition sites that function
synergistically to maximize promoter activity. A second
cis-element binds transcription factors of the ATF/CREB
family (Evans and Scarpulla 1989). The cytochrome c
promoter also contains a recognition site for a transcrip-
tion factor designated nuclear respiratory factor 1, or
NRF-1 (Evans and Scarpulla 1989). NRF-1 is a 68-kD
polypeptide with the presence of a C-terminal transcrip-
tional activation domain comprised of glutamine-con-
taining clusters of hydrophobic amino acid residues
(Chau et al. 1992; Gugneja et al. 1996). Both endogenous
and recombinant proteins bind as a homodimer to pal-
indromic NRF-1 sites through guanine nucleotide con-
tacts over a single turn of the DNA helix (Virbasius et al.
1993a). Serine phosphorylation of the N-terminal do-
main of NRF-1 enhances both its DNA-binding (Gugneja
and Scarpulla 1997) and trans-activation functions (Her-
zig et al. 2000).
NRF-1 has been linked to the transcriptional control of
many genes involved in mitochondrial function and bio-
genesis (Table 1). NRF-1 target genes have been identi-
fied by characterization of functional NRF-1-binding
sites within their promoters. Many NRF-1 target genes
encode subunits of the five respiratory complexes (Vir-
basius et al. 1993a). However, the regulatory network
controlled by NRF-1 extends beyond the respiratory sub-
units to other classes of genes. These include genes in-
volved in assembly of the respiratory apparatus, con-
stituents of the mtDNA transcription and replication
machinery, mitochondrial and cytosolic enzymes of the
heme biosynthetic pathway, and components of mito-
chondrial protein import. Notably, Tfam is an NRF-1
target gene consistent with the postulate that NRF-1
plays an integrative role in nucleo–mitochondrial inter-
actions. This hypothesis has been reinforced by the re-
sults of several recent studies associating increases in
NRF-1 mRNA levels or DNA-binding activity with mi-
tochondrial biogenesis. NRF-1 and its coactivator PGC-1
(see below) are induced as part of the adaptation of skel-
etal muscle to exercise training (Murakami et al. 1998;
Baar et al. 2002). Similar results were obtained in cul-
tured myotubes in response to elevated calcium, which
mimics exercise-induced mitochondrial biogenesis (Ojuka
et al. 2003). Likewise, treatment of rats with a creatine
analog that induces muscle adaptations analogous to
those observed during exercise leads to the activation of
AMP-activated protein kinase and increased NRF-1-DNA
binding activity, cytochrome c content, and mitochon-
drial density (Bergeron et al. 2001). Both NRF-1 and Tfam
mRNAs are elevated in cells depleted of mtDNA, pre-
sumably as a response to increased oxidative stress (Mi-
randa et al. 1999). Lastly, NRF-1 and Tfam are up-regu-
lated in response to lipopolysaccharide-induced oxida-
tive damage to mitochondria, presumably to enhance
mtDNA levels and OXPHOS activity (Suliman et al.
2003).
Perhaps the strongest in vivo link between NRF-1 and
the control of mitochondrial function comes from the
results of targeted disruption of the NRF-1 gene in mice
(Huo and Scarpulla 2001). Homozygosity of the null al-
lele results in lethality between embryonic days 3.5 and
6.5 (E3.5 and E6.5). The null blastocysts fail to grow in
culture despite having a normal morphology. Homozy-
gous null blastocysts are defective in maintaining a mi-
tochondrial membrane potential and have severely re-
duced mtDNA levels. This is not accompanied by in-
creased apoptosis, making it unlikely that the reduction
in mtDNA is associated with a generalized increase in
DNA fragmentation. Moreover, the mature oocytes of
Regulation of mitochondrial biogenesis
GENES & DEVELOPMENT 359
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
heterozygous mothers have a normal complement of
mtDNA, supporting the argument against a defect in
mtDNA amplification during oogenesis. Therefore, the
mtDNA depletion occurs between fertilization and the
blastocyst stage and most likely results from the loss of
a NRF-1-dependent pathway of mtDNA maintenance.
Interestingly, Tfam-null embryos also exhibit severely
depleted levels of mtDNA but survive to E8.5–E10.5
(Larsson et al. 1998). Thus, it is likely that the early
mortality of NRF-1-null embryos results from the com-
bined effects of reduced levels of mtDNA and disruption
of other NRF-1-dependent functions.
Characterization of cytochrome oxidase genes led to
the identification of a second regulatory factor desig-
nated as NRF-2 (for review, see Scarpulla 1997, 1999). A
series of directly repeated NRF-2 sites within the mouse
COXIV promoter overlaps multiple transcription initia-
tion sites and contains additional binding sites for the
ETS-domain family of transcription factors (Virbasius
and Scarpulla 1991; Carter et al. 1992). The complex
binding the NRF-2 sites was purified to homogeneity
from HeLa cell nuclear extracts and is comprised of five
subunits. These include a DNA-binding subunit and
four others (
1
,
2
,
1
, and
2
) that complex with but
alone do not bind DNA. The NRF-2 complexes activate
transcription through four directly repeated ETS-do-
main-binding sites in the COXVb promoter, suggesting
that NRF-2 may also act on multiple respiratory promot-
ers (Virbasius et al. 1993b).
Purification and molecular cloning of all five NRF-2
subunits established that NRF-2 is the human homolog
of mouse GABP (LaMarco and McKnight 1989) and that
the two additional human subunits,
1
and
1
, were mi-
nor splice variants of GABP subunits
1
and
2
(Gugneja
et al. 1995). The function of the non-DNA-binding sub-
units is twofold. First, the GABP
1
subunit, correspond-
ing to NRF-2
1
and NRF-2
2
(Gugneja et al. 1995), has
a dimerization domain that facilitates cooperative bind-
ing of a heterotetrameric complex to tandem binding
sites (Thompson et al. 1991). In solution, GABP exists as
Table 1. Nuclear and mitochondrial genes with NRF-1 and
NRF-2 recognition sites
NRF-1
a
NRF-2
a
Oxidative phosphorylation
Rat cytochrome c +
Human cytochrome c +
Complex I:
Human NADH dehydrogenase subunit
8 (TYKY) +
Complex II:
Human succinate dehydrogenase
subunit B + +
Human succinate dehydrogenase
subunit C + +
Human succinate dehydrogenase
subunit D + +
Complex III:
Human ubiquinone-binding protein +
Human core protein I +
Complex IV:
Rat cytochrome oxidase subunit IV +
Mouse cytochrome oxidase subunit IV +
Mouse cytochrome oxidase subunit Vb + +
Rat cytochrome oxidase subunit Vb + +
Human/primate cytochrome oxidase
subunit Vb + +
Rat cytochrome oxidase subunit VIc +
Human cytochrome oxidase subunit
VIaL + +
Bovine cytochrome oxidase subunit
VIIaL + +
Human cytochrome oxidase subunit
VIIaL +
Bovine cytochrome oxidase subunit
VIIc +
Complex V:
Bovine ATP synthase subunit +
Human ATP synthase c subunit +
Human ATP synthase subunit +
mtDNA transcription and replication
Human Tfam + +
Mouse Tfam +
Rat Tfam +
Mouse MRP RNA +
Human MRP RNA +
Human TFB1 + +
Mouse TFB1 +
Human TFB2 + +
Mouse TFB2 +
HEME biosynthesis
Rat 5-aminolivulinate synthase +
Mouse uroporphyrinogen III synthase + +
Protein import and assembly
Human Tom 20 + +
Human Tom 70 +
Mouse chaperonin 10 +
Human SURF-1 +
Mouse COX17 + +
(continued)
Table 1. (continued)
NRF-1
a
NRF-2
a
Ion channels
Human VDAC3 +
Mouse VDAC3 +
Human VDAC1 +
Shuttles
Human glycerol phosphate
dehydrogenase +
Translation
Human mitochondrial ribosomal S12 + +
a
Original references for the majority of the indicated NRF-1
and/or NRF-2 target genes that are related to mitochondrial
function have been cited elsewhere (Scarpulla 1997, 2002). New
additions include human Tomm 70 (Blesa et al. 2003) and
mouse COX 17 (Takahashi et al. 2002).
Kelly and Scarpulla
360 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
an ␣␤ heterodimer but is induced to form the heterote-
tramer
2
2
by DNA containing two or more binding
sites (Chinenov et al. 2000). The crystal structure of the
heterotetramer bound to DNA has been determined
(Batchelor et al. 1998). The second function of the non-
DNA-binding subunits is to contribute a transcriptional
activation domain. This domain resembles that found in
NRF-1 and has been localized to a region upstream from
the homodimerization domain (Gugneja et al. 1996).
Functional NRF-2 sites have now been identified in
several COX promoters as well as in many other genes
related to respiratory chain expression (for review, see
Scarpulla 2002). As with NRF-1, the list of respiratory
genes containing NRF-2 sites has expanded in recent
years (Table 1). These include genes for Tfam (Larsson
et al. 1998; Rantanen et al. 2001) and the newly discov-
ered TFB factors (Falkenberg et al. 2002; McCulloch et al.
2002) involved in mitochondrial transcription and DNA
replication (Rantanen et al. 2001). Genes encoding three
of the four human succinate dehydrogenase (complex II)
subunits also have both NRF-1 and NRF-2 sites in their
promoters (Au and Scheffler 1998; Elbehti-Green et al.
1998; Hirawake et al. 1999). In many cases, NRF-1 sites
are also present in NRF-2-dependent promoters, but this
is not a general rule. For example, several COX promot-
ers and the rodent Tfam (Choi et al. 2002) and TFB (Ran-
tanen et al. 2003) promoters do not have obvious NRF-1
consensus sites (Table 1). This contrasts with the human
Tfam (Virbasius and Scarpulla 1994) and TFB (R.C.
Scarpulla, unpubl.) promoters, which rely on functional
NRF-1 and NRF-2 recognition sites for their activities.
A subset of respiratory genes does not appear to be
regulated by NRF-1 or NRF-2. Other well-characterized
regulatory factors have been implicated in the expression
of these genes. The transcription factor Sp1 is associated
with the activation and/or repression of cytochrome c
1
(Li et al. 1996b) and adenine nucleotide translocase 2
genes (Li et al. 1996a), both of which lack NRF sites (Zaid
et al. 1999). Sp1 sites are also common to many GC-rich
promoters including those that are NRF-dependent. The
muscle-specific COX subunits, COXVIaH and COXVIII,
are also lacking NRF sites but depend on MEF-2 and/or
E-box consensus elements for their expression (Wan and
Moreadith 1995). Thus, the same or similar factors re-
quired for the expression of other muscle-specific genes
are linked to the regulation of these tissue-specific COX
subunits. In contrast, the promoter of the ubiquitously
expressed liver isoform, COXVIaL, depends on NRF-1
and NRF-2 as well as Sp1 for full activity (Seelan et al.
1996). This is consistent with the observation that in
gene pairs encoding ubiquitous and tissue-specific iso-
forms of a given protein, the NRF-1 site, when present, is
associated with the ubiquitously expressed gene (Virba-
sius et al. 1993a). Finally, the initiator element transcrip-
tion factor YY1 participates in the expression of certain
COX genes. Functional YY1-binding sites have been
detected in the promoters of genes encoding COXVb
(Basu et al. 1997) and COXVIIc (Seelan and Grossman
1997). Multiple YY1 sites in the COXVb promoter bind
YY1 and possibly other factors, and at least one of these
sites helps confer a negative regulatory effect on COXVb
promoter activity (Basu et al. 1997). In the COXVIIc
promoter, two YY1 sites in conjunction with an NRF-2
site act as positive regulators of promoter activity (See-
lan and Grossman 1997). It is also important to note
that regulation of most nuclear genes encoding mito-
chondrial enzymes upstream of the respiratory chain is
NRF-1/NRF-2-independent. For example, genes encod-
ing mitochondrial fatty acid oxidation enzymes are regu-
lated by the peroxisome proliferator-activated receptor
alpha (PPAR) and other NRF-1-independent regulatory
pathways (Gulick et al. 1994). Thus, any unifying tran-
scriptional model of mitochondrial biogenesis needs to
account for the expression of genes that are NRF-inde-
pendent.
There are several reports suggesting that nuclear and
mitochondrial genes are controlled by common cis-act-
ing elements that are the targets of the same or simi-
lar transcription factors. Sequence similarities to the
OXBOX/REBOX (Haraguchi et al. 1994) and Mt (Suzuki
et al. 1995) elements have been localized to the mito-
chondrial D-loop. The ability of these elements and their
nuclear gene counterparts to bind proteins from crude
extracts with the same specificity has been taken as evi-
dence for shared regulatory factors between the two ge-
netic systems (Haraguchi et al. 1994). Similarly, other
nuclear factors, such as thyroid hormone receptors, have
been implicated in mitochondrial gene expression (for
review, see Wrutniak-Cabello et al. 2001). However,
there is no evidence that these proteins can use the mi-
tochondrial transcriptional machinery to direct mito-
chondrial gene expression.
The critical role of transcriptional coactivators
in the mitochondrial biogenic regulatory cascade:
The PPAR coactivator-1 (PGC-1) family
As described above, the mitochondrial biogenic program
involves the integration of multiple transcriptional regu-
latory pathways controlling the expression of both nuclear
and mitochondrial genes. This highlights a mechanistic
enigma fundamental to the control of mitochondrial bio-
genesis. How is the activity of multiple transcription fac-
tors (e.g., NRF-1, NRF-2, PPAR, mtTFA) coordinately
regulated during the mitochondrial biogenic process?
Moreover, in the context of such complex integration,
how is cell- and tissue-specific function achieved? For
example, mitochondria within the brown adipocyte are
poised for uncoupled mitochondrial respiration, whereas
in other tissues such as heart, mitochondrial respiration
is largely coupled for high-level ATP production. To add
to the complexity, skeletal muscle is capable of support-
ing both coupled and uncoupled respiration. New insight
into this problem was provided by the discovery of the
transcriptional coactivator PPAR coactivator 1 (PGC-
1) by Spiegelman and colleagues (Puigserver et al. 1998).
PGC-1
was cloned in a yeast two-hybrid screen for
brown adipose-specific factors that interacted with the
adipogenic nuclear receptor PPAR (Puigserver et al.
1998). PGC-1 serves as a direct transcriptional coacti-
Regulation of mitochondrial biogenesis
GENES & DEVELOPMENT 361
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
vator of PPAR and is a member of a growing list of
proteins that coactivate transcription factors through di-
rect protein–protein interactions (for review, see Knutti
and Kralli 2001; Puigserver and Spiegelman 2003).
Transcriptional coactivators serve multiple functions
including modification of chromatin through posttrans-
lational histone acetylation, direct interaction with the
RNA polymerase II complex, mRNA processing, and re-
cruitment of other transcriptional coactivators (for re-
view, see Robyr et al. 2000; Belandia and Parker 2003).
Present evidence indicates that PGC-1 coactivates its
targets via recruitment of additional coactivators with
histone acetylase activity, such as SRC-1 (Puigserver
et al. 1999). In addition, the PGC-1 molecule contains
domains capable of interacting with and processing pre-
mRNA (Monsalve et al. 2000). PGC-1 also interacts di-
rectly with the TRAP/Mediator complex (Wallberg et al.
2003). Unlike most known transcriptional coactivators,
PGC-1 is unique in that it exhibits a tissue-enriched
expression pattern and is highly inducible (Puigserver
et al. 1998; Knutti and Kralli 2001; Puigserver and Spie-
gelman 2003). PGC-1
is enriched in brown adipose,
heart, slow-twitch skeletal muscle, and kidney—tissues
with high-capacity mitochondrial systems. The expres-
sion of the PGC-1 gene is rapidly induced by cold ex-
posure, short-term exercise, and fasting; physiologic con-
ditions known to increase the demand on mitochondria
to produce heat or ATP (Puigserver et al. 1998; Wu et al.
1999; Goto et al. 2000; Lehman et al. 2000; Baar et al.
2002; Terada et al. 2002; Irrcher et al. 2003; Pilegaard
et al. 2003; Terada and Tabata 2003). These latter obser-
vations suggested that PGC-1 is involved in the physi-
ologic control of mitochondrial function.
Several lines of evidence indicate that the transcrip-
tional coactivator PGC-1 serves as a key regulator of
mitochondrial biogenesis in mammals. First, studies fo-
cused on the biologic function of PGC-1 revealed that it
activates the transcription of mitochondrial uncoupling
protein-1 (UCP-1) through interactions with the nuclear
hormone receptors PPAR and thyroid hormone receptor
(Puigserver et al. 1998). These findings further supported
a role for PGC-1 in the process of mitochondrial un-
coupled respiration and thermogenesis in brown adipose
tissue. Second, forced expression studies in adipogenic
and myogenic mammalian cell lines demonstrated that
PGC-1 markedly induces the expression of NRF-1,
NRF-2, and Tfam (Wu et al. 1999). PGC-1 can also in-
teract directly with and coactivate NRF-1 on the Tfam
gene promoter. Third, studies in primary cardiac myo-
cytes in culture and in the hearts of transgenic mice have
demonstrated that overexpression of PGC-1
up-regu-
lates the expression of genes involved in mitochondrial
fatty acid oxidation, most of which are PPAR targets, in
addition to NRF-1 targets (Lehman et al. 2000). Cardiac-
specific overexpression of PGC-1
in transgenic mice
leads to massive mitochondrial proliferation, ultimately
resulting in cardiomyopathy and death (Lehman et al.
2000). Interestingly, in neonatal cardiac myocytes in cul-
ture, PGC-1 induces mitochondria that support largely
coupled respiration consistent with the known ATP-gen-
erating function of this organelle in heart (Lehman et al.
2000). Lastly, forced expression of PGC-1
in skeletal
muscle of transgenic mice triggers mitochondrial prolif-
eration and the formation of mitochondrial-rich type I,
oxidative (“slow-twitch”) muscle fibers (Lin et al.
2002b). Collectively, these results indicate that PGC-1
is capable of promoting mitochondrial biogenesis through
its coactivating effects on key factors such as NRF-1.
The gain-of-function studies described above provide
compelling evidence that PGC-1 serves as a transcrip-
tional coactivator to promote mitochondrial biogenesis
in postnatal mammalian tissues. Although NRF-1 is a
key target of PGC-1, it is clear that this transcription
factor does not control all of the components of the mi-
tochondrial biogenic response. Multiple PGC-1 targets
have now been identified, indicating that this coactiva-
tor serves as a pleiotropic regulator of multiple pathways
involved in cellular energy metabolism within and out-
side of the mitochondrion (Knutti and Kralli 2001; Puig-
server and Spiegelman 2003). Following the identifica-
tion of PPAR as the initial PGC-1 transcription factor
target, a variety of additional members of the nuclear
receptor superfamily have been shown to interact with
PGC-1. This list includes PPAR (Vega et al. 2000),
thyroid hormone receptor (Puigserver et al. 1998), reti-
noid receptors (Puigserver et al. 1998), glucocorticoid re-
ceptor (Knutti et al. 2000), estrogen receptor (Puigserver
et al. 1998; Knutti et al. 2000; Tcherepanova et al. 2000),
HNF-4 (Rhee et al. 2003), and estrogen-related receptors
(ERRs; Huss et al. 2002; Schreiber et al. 2003). In addi-
tion, several non-nuclear-receptor PGC-1 partners have
been identified, in addition to NRF-1, including myo-
cyte-enhancing factor-2 (MEF-2; Michael et al. 2001) and
FOX-01 (Puigserver et al. 2003). Although several of the
PGC-1 partners serve functions outside of the mito-
chondrion such as HNF-4 and FOX-01 (gluconeogenesis;
Rhee et al. 2003; Puigserver et al. 2003) and MEF-2 (glu-
cose transport; Michael et al. 2001), others are linked to
the mitochondrial biogenic transcriptional regulatory
program. For example, PGC-1 coactivates the nuclear
receptor PPAR, a key regulator of nuclear genes encod-
ing mitochondrial fatty acid oxidation enzymes (Vega
et al. 2000). More recently, PGC-1 was found to coac-
tivate the orphan nuclear receptors ERR and ERR
(Huss et al. 2002; Schreiber et al. 2003). Although the
exact biologic function of ERRs has not been delineated,
ERR and ERR are enriched in tissues with high mito-
chondrial oxidative capacity including brown adipose
tissue and heart. In addition, medium-chain acyl-CoA
dehydrogenase (MCAD), a known PPAR target that
catalyzes the initial step in mitochondrial fatty acid
-oxidation, is also regulated by ERR (Sladek et al.
1997; Vega and Kelly 1997; Huss et al. 2002). These re-
sults suggest that ERR and PPAR may drive distinct
but overlapping mitochondrial pathways downstream of
PGC-1.
PGC-1 is now known to be a member of a family of
transcriptional coactivators. The first PGC-1 relative,
PGC-1-related coactivator (PRC), was identified through
a database search (Andersson and Scarpulla 2001). PRC
Kelly and Scarpulla
362 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
contains several domains that are homologous to PGC-
1 including an acidic N-terminal region, an LXXLL mo-
tif for interacting with nuclear receptors, a proline-rich
region, and regions known to interact with RNA (Fig. 2).
Although overall homology between PGC-1 and PRC is
relatively low, the similarity of domains suggests related
function. In contrast to PGC-1,PRCis largely ubiqui-
tously expressed, is only slightly induced in response to
cold exposure, and is cell-cycle-regulated (Andersson and
Scarpulla 2001). However, functional studies indicate
that PRC may be capable of regulating mitochondrial
function in a manner similar to PGC-1. PRC interacts
directly with and coactivates NRF-1 via natural NRF-1
recognition sites in the 5-ALAS gene promoter (Anders-
son and Scarpulla 2001). Additional experiments have
revealed that PRC activates the transcription of another
known NRF-1 target, cytochrome c, but requires the co-
operation of other factors including CREB (Andersson
and Scarpulla 2001). A third member of the family, PGC-
1
(also termed PGC-1-related estrogen receptor coacti-
vator or PERC), was also identified through database
searching (Kressler et al. 2002; Lin et al. 2002a). PGC-1
exhibits a greater degree of homology to PGC-1 than
PRC (Fig. 2). The expression pattern of PGC-1 exhibits
similarities with that of PGC-1 such as enrichment in
heart and brown adipose tissue. Furthermore, PGC-1
is
induced by fasting but not in response to cold exposure
(Lin et al. 2002a). PGC-1 interacts with HNF-4, NRF-
1, and ERR (Kressler et al. 2002; Lin et al. 2002a; Kamei
et al. 2003). PGC-1 also interacts with Host Cell Factor
(HCF), a cellular protein implicated in cell cycle regula-
tion (Lin et al. 2002a). The relevance of this latter in-
teraction to the regulation of mitochondrial function is
unknown.
The differences in regulation and tissue-expression
patterns of PGC-1 family members suggest that each
confers distinct biologic responses. In support of this
idea, recent work by the Spiegelman laboratory has pro-
vided evidence that PGC-1 and PGC-1 isoforms exert
coactivator-specific bioenergetic effects (St-Pierre et al.
2003). Specifically, overexpression studies in C
2
C
12
myotubes demonstrated that although both PGC-1 and
PGC-1 increase mitochondrial proton leak rates, cells
expressing PGC-1
have a higher proportion of mito-
chondrial respiration linked to proton leak. PGC-1 was
shown to preferentially induce the expression of genes
involved in the removal of reactive oxygen species that
themselves could serve as activators of uncoupling
(St-Pierre et al. 2003).
Upstream signaling events involved in the control
of mitochondrial biogenesis: PGC-1
as an integrative coactivator
The observation that PGC-1
gene expression is rapidly
induced by cold exposure in the brown adipose tissue of
mice spawned a series of observational studies aimed at
determining whether additional physiologic stimuli are
capable of modulating the expression of this coactivator.
It was subsequently shown that PGC-1
gene expression
is induced by exercise in rodent and human skeletal
muscle (Goto et al. 2000; Baar et al. 2002; Terada et al.
2002; Terada and Tabata 2003; Pilegaard et al. 2003) and
by short-term starvation in the heart and liver of mice
(Lehman et al. 2000; Rhee et al. 2003). The transcrip-
tional regulatory mechanisms involved in the regulation
of PGC-1
gene expression in response to physiologic
stimuli are only beginning to be understood. In one ex-
ample, the transcription factor CREB can promote he-
patic gluconeogenesis, in part through its induction of
PGC-1
via direct binding to a functional CRE in the
PGC-1
promoter (Herzig et al. 2001). More recently,
PGC-1
transcription was shown to be regulated by
members of the MEF2 transcription factor family and
repressed by class II histone deacetylases (HDACs; Czu-
bryt et al. 2003).
Signal transduction pathways play a major role in the
physiologic regulation of mitochondrial function and
biogenesis; therefore, it is not surprising that PGC-1
activity and expression are regulated by similar signaling
pathways. In tissues poised for mitochondrial thermo-
genesis, such as brown adipose, the -adrenergic/cAMP
pathway is upstream of the PGC-1-mediated regulation
of targets such as UCP-1 (Puigserver et al. 1998). A sig-
Figure 2. The PGC-1 family of coactiva-
tors. Schematic representation of the pri-
mary structures of PGC-1, PRC, and PGC-
1. The leucine-rich (LXXLL) domains criti-
cal for interaction with nuclear receptors
are also shown. Additional specific shared
domains are indicated as denoted by the key
at the bottom.
Regulation of mitochondrial biogenesis
GENES & DEVELOPMENT 363
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
nificant body of evidence focused largely on skeletal
muscle indicates that in response to contractile activity,
calcium-dependent signaling pathways trigger a cascade
of regulatory events leading to increased formation of
oxidative fiber types and a marked increase in mitochon-
drial number and function (Holloszy and Coyle 1984;
Chin et al. 1998). Several important gain-of-function
studies have now provided evidence for regulatory links
between calcineurin A, calcium/calmodulin-dependent
protein kinase (CaMK), PGC-1, and skeletal muscle mi-
tochondrial biogenesis. First, overexpression of CaMK in
the skeletal muscle of transgenic mice triggers a robust
mitochondrial biogenesis associated with an induction
of PGC-1
expression (Wu et al. 2002). Second, overex-
pression of PGC-1
in the skeletal muscle of transgenic
mice leads to the formation of slow-twitch skeletal
muscle fibers and an induction of genes involved in mi-
tochondrial oxidative metabolism (Lin et al. 2002b).
Third, studies performed in myogenic cell lines indicate
that both calcineurin A and CaMK are capable of acti-
vating PGC-1
gene transcription (Handschin et al.
2003). The calcineurin A-mediated activation of PGC-1
transcription is dependent on MEF2 response elements,
whereas CaMK-mediated regulation requires CREB-bind-
ing sites.
Several other signal transduction pathways have been
implicated in the control of PGC-1 expression and activ-
ity. p38 MAPK activates PGC-1
by releasing repression
of an unidentified factor and by increasing PGC-1 pro-
tein stability (Knutti et al. 2001; Puigserver et al. 2001).
p38 MAPK can also activate the PGC-1 partner,
PPAR, suggesting that activation of this signaling path-
way influences mitochondrial fatty acid oxidation
(Barger et al. 2001). However, the role of the p38 MAPK
pathway in regulating mitochondrial biogenesis is not
known. More recently, evidence has emerged that nitric
oxide (NO) activates mitochondrial biogenesis in a vari-
ety of cell types including adipocytes, and HeLa cells
(Nisoli et al. 2003). The mitochondrial thermogenic re-
sponse is significantly altered in mice lacking eNOS.
This NO effect is dependent on cGMP and linked to
PGC-1
activation. These results raise the intriguing
possibility that mitochondrial biogenesis is one of the
important effects of NO activation. Given the known
role of NO as a vasodilator, it is tempting to speculate
that this key upstream regulatory pathway coordinately
regulates downstream events including an increase in
the capacity to use oxygen in mitochondria.
Summary
Over the past decade, significant new insight has been
gained into the circuitry of molecular regulatory cas-
cades controlling mitochondrial biogenesis and function
(Fig. 3). The interdependence of nuclear and mitochon-
drial genomes has evolved with the emergence of the
mitochondrion as a eukaryotic organelle. It is likely that
the complexity of the mammalian organism mandates a
complex regulatory network that provides for the dy-
namic coordinate control of nuclear and mitochondrial
genes during development and in the adult. This regula-
tory circuitry not only triggers mitochondrial biogenesis
in response to developmental and physiologic cues, but
also confers cell- and tissue-specific features. New in-
sight into the dynamic control of mitochondrial function
and biogenesis has been provided by the identification of
relevant transcription factors, transcriptional coactiva-
tors, and upstream signaling events. However, the
mechanisms involved in the control of cell-specific mi-
tochondrial phenotypes and the full cast of transcrip-
tional regulatory factors comprise an exciting investiga-
tive frontier. New experimental approaches such as the
Figure 3. PGC-1 serves a central integra-
tive role in the transcriptional regulatory
cascade upstream of the mitochondrial bio-
genic response. A schematic representation
of the mitochondrial biogenic regulatory
cascade, including known PGC-1 partners
and putative upstream signaling pathways.
Kelly and Scarpulla
364 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
delineation of tissue-specific mitochondrial proteomes
(Mootha et al. 2003) should provide an excellent frame-
work for future studies aimed at understanding the mo-
lecular events involved in defining the mitochondrial
phenotype.
Acknowledgments
Special thanks to Mary Wingate for assistance with manuscript
preparation and Janice Huss for critical review of the manu-
script. Work in the authors’ laboratories is supported by United
States Public Health Service Grants DK45416, HL58493,
HL57278, HL61006 to D.P.K. and GM32525 to R.C.S. from the
National Institutes of Health.
References
Andersson, U. and Scarpulla, R.C. 2001. PGC-1-Related coacti-
vator, a novel, serum-inducible coactivator of nuclear respi-
ratory factor-1-dependent transcription in mammalian cells.
Mol. Cell. Biol. 21: 3738–3749.
Antoshechkin, I. and Bogenhagen, D.F. 1995. Distinct roles for
two purified factors in transcription of Xenopus mitochon-
drial DNA. Mol. Cell. Biol. 15: 7032–7042.
Au, H.C. and Scheffler, I.E. 1998. Promoter analysis of the hu-
man succinate dehydrogenase iron-protein gene. Both
nuclear respiratory factors NRF-1 and NRF-2 are required.
Eur. J. Biochem. 251: 164–174.
Baar, K., Wende, A.R., Jones, T.E., Marison, M., Nolte, L.A.,
Chen, M., Kelly, D.P., and Holloszy, J.O. 2002. Adaptations
of skeletal muscle to exercise: Rapid increase in the tran-
scriptional coactivator PGC-1. FASEB J. 16: 1879–1886.
Barger, P.M., Browning, A.C., Garner, A.N., and Kelly, D.P.
2001. p38 MAP kinase activates PPAR: A potential role in
the cardiac metabolic stress response. J. Biol. Chem. 276:
44495–44501.
Basu, A., Lenka, N., Mullick, J., and Avadhani, N.G. 1997. Regu-
lation of murine cytochrome oxidase Vb gene expression in
different tissues and during myogenesis—Role of a YY-1 fac-
tor-binding negative enhancer. J. Biol. Chem. 272: 5899–
5908.
Batchelor, A.H., Piper, D.E., De la Brousse, F.C., McKnight, S.L.,
and Wolberger, C. 1998. The structure of GABP/:AnETS
domain ankyrin repeat heterodimer bound to DNA. Science
279: 1037–1041.
Belandia, B. and Parker, M.G. 2003. Nuclear receptors: A ren-
dezvous for chromatin remodeling factors. Cell 114: 277–
280.
Bergeron, R., Ren, J.M., Cadman, K.S., Moore, I.K., Perret, P.,
Pypaert, M., Young, L.H., Semenkovich, C.F., and Shulman,
G.I. 2001. Chronic activation of AMP kinase results in
NRF-1 activation and mitochondrial biogenesis. Am. J.
Physiol. Endocrinol. Metabol. 281: E1340–E1346.
Blesa, J.R, Hernandez, J.M., and Hernandez-Yago, J. 2003. NRF-2
transcription factor is essential in promoting human
Tomm70 gene expression. Mitochondrion (in press).
Bogenhagen, D.F. 1996. Interaction of mtTFB and mtRNA poly-
merase at core promoters for transcription of Xenopus laevis
mtDNA. J. Biol. Chem. 271: 12036–12041.
Bogenhagen, D.F. and Clayton, D.A. 2003. The mitochondrial
DNA replication bubble has not burst. Trends Biochem. Sci.
28: 357–360.
Brown, T.A. and Clayton, D.A. 2002. Release of replication ter-
mination controls mitochondrial DNA copy number after
depletion with 2,3-dideoxycytidine. Nucleic Acids Res.
30: 2004–2010.
Carter, R.S., Bhat, N.K., Basu, A., and Avadhani, N.G. 1992. The
basal promoter elements of murine cytochrome c oxidase
subunit IV gene consist of tandemly duplicated ETS motifs
that bind to GABP-related transcription factors. J. Biol.
Chem. 267: 23418–23426.
Chau, C.A., Evans, M.J., and Scarpulla, R.C. 1992. Nuclear res-
piratory factor 1 activation sites in genes encoding the -sub-
unit of ATP synthase, eukaryotic initiation factor 2, and
tyrosine aminotransferase. Specific interaction of purified
NRF-1 with multiple target genes. J. Biol. Chem. 267: 6999–
7006.
Chin, E.R., Olson, E.N., Richardson, J.A., Yang, Q., Humphries,
C., Shelton, J.M., Wu, H., Zhu, W., Bassel-Duby, R., and
Williams, R.S. 1998. A calcineurin-dependent transcrip-
tional pathway controls skeletal muscle fiber type. Genes &
Dev. 12: 2499–2509.
Chinenov, Y., Henzl, M., and Martin, M.E. 2000. The and
subunits of the GA-binding protein form a stable het-
erodimer in solution. J. Biol. Chem. 275: 7749–7756.
Choi, Y.S., Lee, H.K., and Pak, Y.K. 2002. Characterization of
the 5-flanking region of the rat gene for mitochondrial tran-
scription factor A (Tfam). Biochim. Biophys. Acta Gene
Struct. Expression 1574: 200–204.
Clayton, D.A. 2000. Transcription and replication of mitochon-
drial DNA. Hum. Reprod. 15: 11–17.
Czubryt, M.P., McAnally, J., Fishman, G.I., and Olson, E.N.
2003. Regulation of peroxisome proliferator-activated recep-
tor coactivator 1 (PGC-1) and mitochondrial function by
MEF2 and HDAC5. Proc. Natl. Acad. Sci. 100: 1711–1716.
Dairaghi, D.J., Shadel, G.S., and Clayton, D.A. 1995. Addition
of a 29 residue carboxyl-terminal tail converts a simple
HMG box-containing protein into a transcriptional activa-
tor. J. Mol. Biol. 249: 11–28.
Diffley, J.F. and Stillman, B. 1991. A close relative of the
nuclear, chromosomal high-mobility group protein HMG1
in yeast mitochondria. Proc. Natl. Acad. Sci. 88: 7864–
7868.
Elbehti-Green, A., Au, H.C., Mascarello, J.T., Ream-Robinson,
D., and Scheffler, I.E. 1998. Characterization of the human
SDHC gene encoding one of the integral membrane proteins
of succinate-quinone oxidoreductase in mitochondria. Gene
213: 133–140.
Evans, M.J. and Scarpulla, R.C. 1988. Both upstream and intron
sequence elements are required for elevated expression of
the rat somatic cytochrome c gene in COS-1 cells. Mol. Cell.
Biol. 8: 35–41.
———. 1989. Interaction of nuclear factors with multiple sites
in the somatic cytochrome c promoter. Characterization of
upstream NRF-1, ATF and intron Sp1 recognition sites.
J. Biol. Chem. 264: 14361–14368.
Falkenberg, M., Gaspari, M., Rantanen, A., Trifunovic, A., Lars-
son, N.-G., and Gustafsson, C.M. 2002. Mitochondrial tran-
scription factors B1 and B2 activate transcription of human
mtDNA. Nat. Genet. 31: 289–294.
Fisher, R.P., Lisowsky, T., Parisi, M.A., and Clayton, D.A. 1992.
DNA wrapping and bending by a mitochondrial high mobil-
ity group-like transcriptional activator protein. J. Biol.
Chem. 267: 3358–3367.
Garesse, R. and Vallejo, C.G. 2001. Animal mitochondrial bio-
genesis and function: A regulatory cross-talk between two
genomes. Gene 263: 1–16.
Goto, M., Terada, S., Kato, M., Katoh, M., Yokozeki, T., Tabata,
I., and Shimokawa, T. 2000. cDNA cloning and mRNA
analysis of PGC-1 in epitrochlearis muscle in swimming-
Regulation of mitochondrial biogenesis
GENES & DEVELOPMENT 365
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
exercised rats. Biochem. Biophys. Res. Commun. 274: 350–
354.
Gugneja, S. and Scarpulla, R.C. 1997. Serine phosphorylation
within a concise amino-terminal domain in nuclear respira-
tory factor 1 enhances DNA binding. J. Biol. Chem. 272:
18732–18739.
Gugneja, S., Virbasius, J.V., and Scarpulla, R.C. 1995. Four struc-
turally distinct, non-DNA-binding subunits of human
nuclear respiratory factor 2 share a conserved transcriptional
activation domain. Mol. Cell. Biol. 15: 102–111.
Gugneja, S., Virbasius, C.A., and Scarpulla, R.C. 1996. Nuclear
respiratory factors 1 and 2 utilize similar glutamine-contain-
ing clusters of hydrophobic residues to activate transcrip-
tion. Mol. Cell. Biol. 16: 5708–5716.
Gulick, T., Cresci, S., Caira, T., Moore, D.D., and Kelly, D.P.
1994. The peroxisome proliferator activated receptor regu-
lates mitochondrial fatty acid oxidative enzyme gene expres-
sion. Proc. Natl. Acad. Sci. 91: 11012–11016.
Handschin, C., Rhee, J., Lin, J., Tam, P.T., and Spiegelman, B.M.
2003. An autoregulatory loop controls peroxisome prolifera-
tor-activated receptor coactivator 1 expression in muscle.
Proc. Natl. Acad. Sci. 100: 7111–7116.
Haraguchi, Y., Chung, A.B., Neill, S., and Wallace, D.C. 1994.
OXBOX and REBOX, overlapping promoter elements of the
mitochondrial F
0F1-ATP
synthase subunit gene. OXBOX/
REBOX in the ATPsyn promoter. J. Biol. Chem. 269: 9330–
9334.
Herzig, R.P., Scacco, S., and Scarpulla, R.C. 2000. Sequential
serum-dependent activation of CREB and NRF-1 leads to
enhanced mitochondrial respiration through the induction
of cytochrome c. J. Biol. Chem. 275: 13134–13141.
Herzig, S., Long, F., Jhala, U.S., Hedrick, S., Quinn, R., Bauer, A.,
Rudolph, D., Schutz, G., Yoon, C., Puigserver, P., et al. 2001.
CREB regulates hepatic gluconeogenesis through the coacti-
vator PGC-1. Nature 413: 179–183.
Hirawake, H., Taniwaki, M., Tamura, A., Amino, H., Tomit-
suka, E., and Kita, K. 1999. Characterization of the human
SDHD gene encoding the small subunit of cytochrome b
(cybS) in mitochondrial succinate-ubiquinone oxidoreduc-
tase. Biochim. Biophys. Acta 1412: 295–300.
Holloszy, J.O. and Coyle, E.F. 1984. Adaptations of skeletal
muscle to endurance exercise and their metabolic conse-
quences. J. Applied Physiol. 56: 831–838.
Huo, L. and Scarpulla, R.C. 2001. Mitochondrial DNA instabil-
ity and peri-implantation lethality associated with targeted
disruption of nuclear respiratory factor 1 in mice. Mol. Cell.
Biol. 21: 644–654.
Huss, J.M., Kopp, R.P., and Kelly, D.P. 2002. PGC-1 coacti-
vates the cardiac-enriched nuclear receptors estrogen-related
receptor- and -. J. Biol. Chem. 277: 40265–40274.
Irrcher, I., Adhihetty, P.J., Sheehan, T., Joseph, A.-M., and Hood,
D.A. 2003. PPAR coactivator-1 expression during thyroid
hormone- and contractile activity-induced mitochondrial
adaptations. Am. J. Physiol. Cell Physiol. 284: C1669–
C1677.
Kamei, Y., Ohizumi, H., Fujitani, Y., Nemoto, T., Tanaka, T.,
Takahashi, N., Kawada, T., Miyoshi, M., Ezaki, O., and Kaki-
zuka, A. 2003. PPAR coactivator 1/ERR ligand 1 is an ERR
protein ligand, whose expression induces a high-energy
expenditure and antagonizes obesity. Proc. Natl. Acad. Sci.
100: 12378–12383.
Knutti, D. and Kralli, A. 2001. PGC-1, a versatile coactivator.
Trends Endocrinol. Metab. 12: 360–365.
Knutti, D., Kaul, A., and Kralli, A. 2000. A tissue-specific co-
activator of steroid receptors. Mol. Cell. Biol. 20: 2411–
2422.
Knutti, D., Kressler, D., and Kralli, A. 2001. Regulation of the
transcriptional coactivator PGC-1 via MAPK-sensitive inter-
action with a corepressor. Proc. Natl. Acad. Sci. 98: 9713–
9718.
Kressler, D., Schreiber, S.N., Knutti, D., and Kralli, A. 2002. The
PGC-1-related protein PERC is a selective coactivator of es-
trogen receptor . J. Biol. Chem. 277: 13918–13925.
LaMarco, K.L. and McKnight, S.L. 1989. Purification of a set of
cellular polypeptides that bind to the purine-rich cis-regula-
tory element of herpes simplex virus immediate early genes.
Genes & Dev. 3: 1372–1383.
Larsson, N.-G., Oldfors, A., Holme, E., and Clayton, D.A. 1994.
Low levels of mitochondrial transcription factor A in mito-
chondrial DNA depletion. Biochem. Biophys. Res. Com-
mun. 200: 1374–1381.
Larsson, N.-G., Wang, J.M., Wilhelmsson, H., Oldfors, A., Rus-
tin, P., Lewandoski, M., Barsh, G.S., and Clayton, D.A. 1998.
Mitochondrial transcription factor A is necessary for
mtDNA maintenance and embryogenesis in mice. Nat.
Genet. 18: 231–236.
Lehman, J.J., Barger, P.M., Kovacs, A., Saffitz, J.E., Medeiros, D.,
and Kelly, D.P. 2000. PPAR coactivator-1 (PGC-1) pro-
motes cardiac mitochondrial biogenesis. J. Clin. Invest.
106: 847–856.
Li, R., Hodny, Z., Luciakova, K., Barath, P., and Nelson, B.D.
1996a. Sp1 activates and inhibits transcription from separate
elements in the proximal promoter of the human adenine
nucleotide translocase 2 (ANT2) gene. J. Biol. Chem. 271:
18925–18930.
Li, R., Luciakova, K., and Nelson, B.D. 1996b. Expression of the
human cytochrome c
1
gene is controlled through multiple
Sp1-binding sites and an initiator region. Eur. J. Biochem.
241: 649–656.
Lin, J., Puigserver, P., Donovan, J., Tarr, P., and Spiegelman,
B.M. 2002a. Peroxisome proliferator-activated receptor co-
activator 1 (PGC-1), a novel PGC-1-related transcription
coactivator associated with host cell factor. J. Biol. Chem.
277: 1645–1648.
Lin, J., Wu, H., Tarr, P.T., Zhang, C.Y., Wu, Z., Boss, O., Mi-
chael, L.F., Puigserver, P., Isotanni, E., Olson, E.N., et al.
2002b. Transcriptional co-activator PGC-1 drives the
formation of slow-twitch muscle fibres. Nature 418: 797–
801.
McCulloch, V. and Shadel, G.S. 2003. Human mitochondrial
transcription factor B1 interacts with the C-terminal activa-
tion region of h-mtTFA and stimulates transcription inde-
pendently of its RNA methyltransferase activity. Mol. Cell.
Biol. 23: 5816–5824.
McCulloch, V., Seidel-Rogol, B.L., and Shadel, G.S. 2002. A hu-
man mitochondrial transcription factor is related to RNA
adenine methyltransferases and binds S-adenosylmethio-
nine. Mol. Cell. Biol. 22: 1116–1125.
Michael, L.F., Wu, Z., Cheatham, R.B., Puigserver, P., Adel-
mant, G., Lehman, J.J., Kelly, D.P., and Spiegelman, B.M.
2001. Restoration of insulin-sensitive glucose transporter
(GLUT4) gene expression in muscle cells by the transcrip-
tional coactivator PGC-1. Proc. Natl. Acad. Sci. 98: 3820–
3825.
Miranda, S., Foncea, R., Guerreo, J., and Leighton, F. 1999. Oxi-
dative stress and upregulation of mitochondrial biogenesis
genes in mitochondrial DNA-depleted HeLa cells. Biochem.
Biophys. Res. Commun. 258: 44–49.
Monsalve, M., Wu, Z., Adelmant, G., Puigserver, P., Fan, M.,
and Spiegelman, B.M. 2000. Direct coupling of transcription
and mRNA processing through the thermogenic coactivator
PGC-1. Mol. Cell 6: 307–316.
Kelly and Scarpulla
366 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
Mootha, V.K., Bunkenborg, J., Olsen, J.V., Hjerrid, M., Wisniew-
ski, J.R., Stahl, E., Bolouri, M.S., Ray, H.N., Sihag, S., Kamal,
M., et al. 2003. Integrated analysis of protein composition,
tissue diversity, and gene regulation in mouse mitochondria.
Cell 115: 629–640.
Murakami, T., Shimomura, Y., Yosimura, A., Sokabe, M., and
Fujitsuka, N. 1998. Induction of nuclear respiratory factor-1
expression by an acute bout of exercise in rat muscle. Bio-
chem. Biophys. Acta 1381: 113–122.
Nisoli, E., Clementi, E., Paolucci, C., Cozzi, V., Tonello, C.,
Sciorati, C., Bracale, R., Valerio, A., Francolini, M.,
Moncada, S., et al. 2003. Mitochondrial biogenesis in mam-
mals: The role of endogenous nitric oxide. Science 299: 896–
899.
Ojuka, E.O., Jones, T.E., Han, D.H., Chen, M., and Holloszy,
J.O. 2003. Raising Ca
2+
in L6 myotubes mimics effects of
exercise on mitochondrial biogenesis in muscle. FASEB J.
17: 675–681.
Parisi, M.A., Xu, B., and Clayton, D.A. 1993. A human mito-
chondrial transcriptional activator can functionally replace a
yeast mitochondrial HMG-box protein both in vivo and in
vitro. Mol. Cell. Biol. 13: 1951–1961.
Pilegaard, H., Saltin, B., and Neufer, P.D. 2003. Exercise induces
transient transcriptional activation of the PGC-1 gene in
human skeletal muscle. J. Physiol. 546: 851–858.
Poulton, J., Morten, K., Freeman-Emmerson, C., Potter, C.,
Sewry, C., Dubowitz, V., Kidd, H., Stephenson, J., White-
house, W., Hansen, F.J., et al. 1994. Deficiency of the human
mitochondrial transcription factor h-mtTFA in infantile mi-
tochondrial myopathy is associated with mtDNA depletion.
Hum. Mol. Genet. 3: 1763–1769.
Poyton, R.O. and McEwen, J.E. 1996. Crosstalk between nuclear
and mitochondrial genomes. Annu. Rev. Biochem. 65: 563–
607.
Puigserver, P. and Spiegelman, B.M. 2003. Peroxisome prolifera-
tor-activated receptor- coactivator 1 (PGC-1): Transcrip-
tional coactivator and metabolic regulator. Endocrine Rev.
24: 78–90.
Puigserver, P., Wu, Z., Park, C.W., Graves, R., Wright, M., and
Spiegelman, B.M. 1998. A cold-inducible coactivator of
nuclear receptors linked to adaptive thermogenesis. Cell
92: 829–839.
Puigserver, P., Adelmant, G., Wu, Z., Fan, M., Xu, J., O’Malley,
B., and Spiegelman, B.M. 1999. Activation of PPAR coacti-
vator-1 through transcription factor docking. Science 286:
1368–1371.
Puigserver, P., Rhee, J., Lin, J., Wu, Z., Yoon, J.C., Zhang, C.-Y.,
Kraquss, S., Mootha, V.K., Lowell, B.B., and Spiegelman,
B.M. 2001. Cytokine stimulation of energy expenditure
through p38 MAP kinase activation of PPAR coactivator-1.
Mol. Cell 8: 971–982.
Puigserver, P., Rhee, J., Donovan, J., Walkey, C.J., Yoon, J.C.,
Oriente, F., Kitamura, Y., Altomonte, J., Dong, H., Accili, D.,
et al. 2003. Insulin-regulated hepatic gluconeogenesis
through FOXO1–PGC-1 interaction. Nature 423: 550–
555.
Rantanen, A., Jansson, M., Oldfors, A., and Larsson, N.-G. 2001.
Downregulation of Tfam and mtDNA copy number during
mammalian spermatogenesis. Mamm. Genome 12: 787–
792.
Rantanen, A., Gaspari, M., Falkenberg, M., Gustafssn, C.M.,
and Larsson, N.-G. 2003. Characterization of the mouse
genes for mitochondrial transcription factors B1 and B2.
Mamm. Genome 14: 1–6.
Rhee, J., Inoue, Y., Yoon, J.C., Puigserver, P., Fam, M., Gonzalez,
F.J., and Spiegelman, B.M. 2003. Regulation of hepatic fast-
ing response by PPAR coactivator-1 (PGC-1): Require-
ment for hepatocyte nuclear factor 4 in gluconeogenesis.
Proc. Natl. Acad. Sci. 100: 4012–4017.
Robyr, D., Wolffe, A.P., and Wahli, W. 2000. Nuclear hormone
receptor coregulators in action: Diversity for shared tasks.
Mol. Endocrinol. 14: 329–347.
Scarpulla, R.C. 1997. Nuclear control of respiratory chain ex-
pression in mammalian cells. J. Bioenerg. Biomembr. 29:
109–119.
———. 1999. Nuclear transcription factors in cytochrome c and
cytochrome oxidase expression. In Frontiers of cellular bio-
energetics: Molecular biology, biochemistry, and physiopa-
thology (ed. S. Papa et al.), pp. 553–591. Plenum, London.
———. 2002. Nuclear activators and coactivators in mamma-
lian mitochondrial biogenesis. Biochim. Biophys. Acta 1576:
1–14.
Schreiber, S.N., Knutti, D., Brogli, K., Uhlmann, T., and Kralli,
A. 2003. The transcriptional coactivator PGC-1 regulates the
expression and activity of the orphan nuclear receptor estro-
gen-related receptor (ERR). J. Biol. Chem. 278: 9013–
9018.
Seelan, R.S. and Grossman, L.I. 1997. Structural organization
and promoter analysis of the bovine cytochrome c oxidase
subunit VIIc gene—A functional role for YY1. J. Biol. Chem.
272: 10175–10181.
Seelan, R.S., Gopalakrishnan, L., Scarpulla, R.C., and Gross-
man, L.I. 1996. Cytochrome c oxidase subunit VIIa liver
isoform: Characterization and identification of promoter
elements in the bovine gene. J. Biol. Chem. 271: 2112–
2120.
Shadel, G.S. and Clayton, D.A. 1997. Mitochondrial DNA
maintenance in vertebrates. Annu. Rev. Biochem. 66: 409–
435.
Sladek, R., Bader, J.-A., and Giguere, V. 1997. The orphan
nuclear receptor estrogen-related receptor is a transcrip-
tional regulator of the human medium-chain acyl coenzyme
a dehydrogenase gene. Mol. Cell. Biol. 17: 5400–5409.
St-Pierre, J., Lin, J., Krauss, S., Tarr, P.T., Yang, R., Newgard,
C.B., and Spiegelman, B.M. 2003. Bioenergetic analysis of
peroxisome proliferator-activated receptor coactivators 1
and 1 (PGC-1 and PGC-1) in muscle cells. J. Biol. Chem.
278: 26597–26603.
Suliman, H.B., Carraway, M.S., Welty-Wolf, K.E., Whorton,
A.R., and Piantados, C.A. 2003. Lipopolysaccharide stimu-
lates mitochondrial biogenesis via activation of nuclear res-
piratory factor-1. J. Biol. Chem. 278: 41510–41518.
Suzuki, H., Suzuki, S., Kumar, S., and Ozawa, T. 1995. Human
nuclear and mitochondrial Mt element-binding proteins to
regulatory regions of the nuclear respiratory genes and to the
mitochondrial promoter region. Biochem. Biophys. Res.
Commun. 213: 204–210.
Takahashi, Y., Kako, K., Arai, H., Ohishi, T., Inada, Y. Take-
hara, A., Fukamizu, A., and Munekata, E. 2002. Character-
ization and identification of promoter elements in the
mouse COX17 gene. Biochim. Biophys. Acta 1574: 359–
364.
Tcherepanova, I., Puigserver, P., Norris, J.D., Speigelman, B.M.,
and McDonnell, D.P. 2000. Modulation of estrogen recep-
tor- transcriptional activity by coactivator PGC-1. J. Biol.
Chem. 275: 16302–16308.
Terada, S. and Tabata, I. 2003. Effects of acute bouts of running
and swimming exercise on PGC-1 protein expression in rat
epitrochlearis and soleus muscle. Am. J. Physiol. Endocrinol.
Metabol. 286: E208–E216.
Terada, S., Goto, M., Kato, M., Kawanaka, K., Shimokawa, T.,
and Tabata, I. 2002. Effects of low-intensity prolonged exer-
Regulation of mitochondrial biogenesis
GENES & DEVELOPMENT 367
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
cise on PGC-1 mRNA expression in rat epitrochlearis
muscle. Biochem. Biophys. Res. Commun. 296: 350–354.
Thompson, C.C., Brown, T.A., and McKnight, S.L. 1991. Con-
vergence of ETS- and notch-related structural motifs in a
heteromeric DNA binding complex. Science 253: 762–768.
Tiranti, V., Savoia, A., Forti, F., D’Apolito, M.F., Centra, M.,
Racchi, M., and Zeviani, M. 1997. Identification of the gene
encoding the human mitochondrial RNA polymerase (h-
mtRPOL) by cyberscreening of the expressed sequence tags
database. Hum. Mol. Genet. 6: 615–625.
Vega, R.B. and Kelly, D.P. 1997. A role for estrogen-related re-
ceptor in the control of mitochondrial fatty acid -oxida-
tion during brown adipocyte differentiation. J. Biol. Chem.
272: 31693–31699.
Vega, R.B., Huss, J.M., and Kelly, D.P. 2000. The coactivator
PGC-1 cooperates with peroxisome proliferator-activated re-
ceptor in transcriptional control of nuclear genes encoding
mitochondrial fatty acid oxidation enzymes. Mol. Cell. Biol.
20: 1868–1876.
Virbasius, J.V. and Scarpulla, R.C. 1991. Transcriptional activa-
tion through ETS domain binding sites in the cytochrome c
oxidase subunit IV gene. Mol. Cell. Biol. 11: 5631–5638.
———. 1994. Activation of the human mitochondrial transcrip-
tion factor A gene by nuclear respiratory factors: A potential
regulatory link between nuclear and mitochondrial gene ex-
pression in organelle biogenesis. Proc. Natl. Acad. Sci. 91:
1309–1313.
Virbasius, C.A., Virbasius, J.V., and Scarpulla, R.C. 1993a.
NRF-1, an activator involved in nuclear-mitochondrial in-
teractions, utilizes a new DNA-binding domain conserved in
a family of developmental regulators. Genes & Dev. 7: 2431–
2445.
Virbasius, J.V., Virbasius, C.A., and Scarpulla, R.C. 1993b. Iden-
tity of GABP with NRF-2, a multisubunit activator of cyto-
chrome oxidase expression, reveals a cellular role for an ETS
domain activator of viral promoters. Genes & Dev. 7: 380–
392.
Wallberg, A.E., Yamamura, S., Malik, S., Spiegelman, B.M., and
Roeder, R.G. 2003. Coordination of p300-mediated chroma-
tin remodeling and TRAP/mediatro function through coac-
tivator PGC-1. Mol. Cell 12: 1137–1149.
Wan, B. and Moreadith, R.W. 1995. Structural characterization
and regulatory element analysis of the heart isoform of cy-
tochrome c oxidase VIa. J. Biol. Chem. 270: 26433–26440.
Wrutniak-Cabello, C., Casas, F., and Cabello, G. 2001. Thyroid
hormone action in mitochondria. J. Mol. Endocrinol. 26: 67–
77.
Wu, Z., Puigserver, P., Andersson, U., Zhang, C., Adelmant, G.,
Mootha, V., Troy, A., Cinti, S., Lowell, B., Scarpulla, R.C.,
et al. 1999. Mechanisms controlling mitochondrial biogen-
esis and respiration through the thermogenic coactivator
PGC-1. Cell 98: 115–124.
Wu, H., Kanatous, S.B., Thurmond, F.A., Gallardo, T., Isotani,
E., Bassel-Duby, R., and Williams, R.S. 2002. Regulation of
mitochondrial biogenesis in skeletal muscle by CaMK. Sci-
ence 296: 349–352.
Zaid, A., Li, R., Luciakova, K., Barath, P., Nery, S., and Nelson,
B.D. 1999. On the role of the general transcription factor Sp1
in the activation and repression of diverse mammalian oxi-
dative phosphorylation genes. J. Bioenerg. Biomembr. 31:
129–135.
Zitomer, R.S. and Lowry, C.V. 1992. Regulation of gene expres-
sion by oxygen in Saccharomyces cerevisiae. Microbiol.
Rev. 56: 1–11.
Kelly and Scarpulla
368 GENES & DEVELOPMENT
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from
10.1101/gad.1177604Access the most recent version at doi:
18:2004, Genes Dev.
Daniel P. Kelly and Richard C. Scarpulla
biogenesis and function
Transcriptional regulatory circuits controlling mitochondrial
References
http://genesdev.cshlp.org/content/18/4/357.full.html#ref-list-1
This article cites 103 articles, 54 of which can be accessed free at:
License
Service
Email Alerting
click here.right corner of the article or
Receive free email alerts when new articles cite this article - sign up in the box at the top
Cold Spring Harbor Laboratory Press
Cold Spring Harbor Laboratory Press on September 1, 2024 - Published by genesdev.cshlp.orgDownloaded from