DataPDF Available

Document S2. Article plus Supplemental Information

Authors:
Stem Cell Reports
Ar ticle
The Satellite Cell Niche Regulates the Balance between Myoblast
Differentiation and Self-Renewal via p53
Valentina Flamini,
1
Rachel S. Ghadiali,
1
Philipp Antczak,
2,3
Amy Rothwell,
1
Jeremy E. Turnbull,
1
and Addolorata Pisconti
1,
*
1
Department of Biochemistry, Institute of Integrative Biology, University of Liverpool, Liverpool L69 7ZB, UK
2
Department of Functional Genomics, Institute of Integrative Biology, University of Liverpool, Liverpool L69 7ZB, UK
3
Computational Biology Facility, Institute of Integrative Biology, University of Liverpool, Liverpool L69 7ZB, UK
*Correspondence: pisconti@liverpool.ac.uk
https://doi.org/10.1016/j.stemcr.2018.01.007
SUMMARY
Satellite cells are adult muscle stem cells residing in a specialized niche that regulates their homeostasis. How niche-generated sig-
nals integrate to regulate gene expression in satellite cell-derived myoblasts is poorly understood. We undertook an unbiased
approach to study the effect of the satellite cell niche on satellite cell-derived myoblast transcriptional regulation and identified
the tumor suppressor p53 as a key player in the regulation of myoblast quiescence. After activation and proliferation, a subpopu-
lation of myoblasts cultured in the presence of the niche upregulates p53 and fails to differentiate. When satellite cell self-renewal
is modeled ex vivo in a reserve cell assay, myoblasts treated with Nutlin-3, which increases p53 levels in the cell, fail to differ-
entiate and instead become quiescent. Since both these Nutlin-3 effects are rescued by small interfering RNA-mediated p53
knockdown, we conclude that a tight control of p53 levels in myoblasts regulates the balance between differentiation and return
to quiescence.
INTRODUCTION
Satellite cells (SCs) are quiescent muscle stem cells
residing in a specialized anatomical niche located be-
tween the plasma membrane of the muscle fiber and
the surrounding basal lamina (Mashinchian et al.,
2017). In response to muscle damage, SCs become acti-
vated and re-enter the cell cycle. After one or more
rounds of proliferation, the vast majority of SC-derived
muscle progenitors (called myoblasts) exit the cell cycle
and enter a terminal G0 phase that leads to differentia-
tion, followed by fusion to existing damaged muscle
fibers to repair them or one-another to generate new
muscle fibers. During this process, a small portion of
myoblasts do not differentiate and rather enters a revers-
ible G0 phase of the cell cycle, effectively replenishing
the pool of quiescent SCs (Olguı
´n and Pisconti, 2012).
The mechanisms that regulate these fate decisions have
been investigated, and at least two models have been pro-
posed. In the first model, activated SCs divide asymmet-
rically upon activation giving rise to a daughter cell
that self-renews and another daughter cell that becomes
a myoblast and gives rise to a myogenic progeny (Du-
mont et al., 2015; Kuang et al., 2007; Troy et al., 2012).
In a second model, proliferating myoblasts are induced
to overexpress Pax7, which in turn inhibits myogenin
expression and promotes entry into a mitotically quies-
cent state (Olguı
´n and Olwin, 2004; Wen et al., 2012).
In both cases, a key role appears to be played by the extra-
cellular environment, called the SC niche (Mashinchian
et al., 2017).
The SC niche is the complex set of molecules surround-
ing the SC in its anatomical location and the receptors
that are expressed on its surface. Several of these molecules
play important roles in driving SC fate (Mashinchian et al.,
2017). However, it is also well established that when SCs
are completely stripped of their native niche and cultured
on a proteinaceous substrate, usually collagen, laminin,
or gelatin, they retain the capacity to recapitulate the
fate choices normally made in the presence of the niche,
including proliferation, differentiation, fusion, and gener-
ation of a population of quiescent cells resembling self-
renewed SCs (Olguı
´n and Olwin, 2004). These observations
raise two questions: (1) Are SCs primed to follow the
myogenic program regardless of the presence of the niche?
(2) Is the transcriptional program that drives these fate
decisions in SCs the same in the presence of the niche
and its absence?
Here we attempt to answer these questions by investi-
gating gene expression in SC-derived myoblasts cultured
under two different conditions: in the presence of their
native niche (on isolated myofibers) or in its absence (on
gelatin-coated dishes). We show how myoblast gene expres-
sion is affected by the presence of the niche and identify the
p53 gene network as a key regulator of myoblast fate in the
presence of the SC niche. Lastly, we show that a sustained
increase in p53 during myoblast cell-cycle exit inhibits
myoblast differentiation while promoting quiescence.
970 Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018 jª2018 The Authors.
This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).
RESULTS
Dispersed and Myofiber-Associated Myoblasts Exit the
Cell Cycle and Initiate Differentiation with Similar
Timing
When SCs are isolated from the muscle tissue and cultured
on gelatin-coated dishes, they extensively proliferate for
the first 2–3 days in culture (Figures 1A, 1B, and S1). Prolifer-
ating SCs express Pax7,Myf5,andMyoD1 and are often
referred to as myoblasts. On the fourth day in culture, a few
myotubes can be already observed (Figure S1). Indeed, myo-
genin-positive (MYOG+) cells are occasionally observed on
the third day in culture (Figure 1B), suggesting that SC-
derived myoblasts in dispersed cultures begin to exit the
cell cycle and undergo terminal differentiation between
48 and 72 hr after isolation. Similarly, on day 3 in culture,
MYOG+ cells are observed amongst myofiber-associated
myoblasts (Figures 1C and 1D), which are cultured in the
same medium as dispersed myoblasts. This suggests that
the timing ofmyoblast cell-cycle exit and entryinto terminal
differentiation are comparable regardless of the presence of
the niche. To test whether these comparable timings were
driven by comparable transcriptional programs, we carried
out a globalgene expression analysis of SC-derived myoblasts
cultured either in dispersed cultures or on explanted myofib-
ers. We profiled gene expression in myoblasts from both cell
culture types at 48 and 72 hr after isolation, when cell-cycle
exit and commitment to terminal differentiation appear to
occur under both culture conditions (Figures 1A–1D).
Myoblast Cell-Cycle Exit Is Associated with Different
Transcriptional Signatures in the Presence or Absence
of the SC Niche
We collected four biological replicates for each time point
(48 and 72 hr) in each culture condition and analyzed
gene expression by microarray technology. The extent of
reproducibility across replicates was excellent (Figures S2A
and S2B). By contrast, the myoblast transcriptome at 48 hr
was remarkably different from the transcriptome at 72 hr
under both culture conditions, as evidenced by the large
Figure 1. Cell-Cycle Exit and Terminal
Differentiation Are Induced in Both Myo-
fiber-Associated and Dispersed Myoblasts
between 48 and 72 hr after Isolation
(A and B) Dispersed myoblasts cultured
on gelatin-coated plates show a rounded
morphology (A) and proliferate extensively
in the first 2–3 days as revealed by positive
staining for the cell-cycle marker KI67+.
No differentiating cells are detected at 48 hr
after isolation (B). As early as 72 hr post-
isolation occasionally MYOG+ cells are de-
tected in dispersed cultures (B), arrow.
(C and D) For the first 2 days myofiber-
associated myoblasts (C) proliferate as re-
vealed by positive staining for KI67+ and
absence of differentiating (MYOG+) cells
(D). At 72 hr after isolation a few MYOG+
cells are occasionally detected (D), arrow.
(E and F) Genes differentially expressed
between 48 and 72 hr in dispersed (E)
and myofiber-associated (F) myoblasts were
mapped to canonical gene networks using
IPA, revealing that the top most enriched
gene network in dispersed myoblasts is
centered around Erk1/2 downregulation (E),
while the top most enriched network in
myofiber-associated myoblasts is centered
around Trp53 upregulation (F). Genes labeled
in green are downregulated, genes labeled in
red are upregulated at 72 hr compared to
48 hr. The color intensity is proportional to
the extent of up- or downregulation.
Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018 971
number of differentially expressed genes (at q < 0.01) de-
tected between 48 and 72 hr undereither culture conditions:
1,810 in dispersed myoblasts and 1,999 in myofiber-associ-
ated myoblasts. Interestingly, when we compared the 72 hr/
48 hr fold changes between the two culture conditions,
it appeared evident that gene expression changes between
48 and 72 hr were different in the two culture conditions
(Figure S2C). To gain insightinto the molecular mechanisms
that were associated with these dramatic changes in the
transcriptional signature of myoblasts between 48 and
72 hr in either dispersed or myofiber-associated cultures,
we mapped the differentially expressed genes to known
gene networks using Ingenuity Pathway Analysis (IPA).
The top most enriched network to which differentially
expressed genes from dispersed myoblasts mapped, was
centered around a decrease in the intracellular kinases Erk1
and Erk2 (Figure 1E). In contrast, the top most enriched
network to which differentially expressed genes from myo-
fiber-associated myoblasts mapped, was centered around
an increase in the tumor suppressor Trp53 (p53) (Figure 1F).
ERK1/2 are key promoters of myoblast proliferation (Jones
et al., 2001) and, similarly, an increase in p53 levels is
expected to lead to cell-cycle arrest (Levine, 1997). Thus,
these results are consistent with our initial hypothesis
that between 48 and 72 hr both dispersed and myofiber-
associated myoblasts prepare to exit the cell cycle, though
via different molecular mechanisms.
The Signaling Pathways that Regulate Cell-Cycle Exit
in the Presence or Absence of the Niche Are Different
To further our understanding of the molecular mechanisms
regulating SC gene expression in the presence and absence
of the SC niche, we analyzed the canonical signaling path-
ways that were enriched between 48 and 72 hr in dispersed
and myofiber-associated myoblasts using IPA, which as-
signs an activation score to a signaling pathway based on
the direction and extent of change of the differentially
expressed genes mapped to that signaling pathway. The
first observation was that the vast majority of canonical
signaling pathways moved in different directions between
48 and 72 hr in myofiber-associated versus dispersed myo-
blasts (Figure 2A).
When we then analyzed in more detail the IPA category
Cell Cycle, the most striking feature was a strong activation
of Cdk5 signaling (Figure 2B) observed in both culture condi-
tions, indicating cell-cycle arrest and preparation to differ-
entiate (Lazaro et al., 1997; Sarker and Lee, 2004). In myo-
fiber-associated myoblasts this was accompanied by an
even stronger activation of ATM signaling, whichis upstream
of p53 and also induces cell-cycle arrest (Figure 2B). Consis-
tently, signaling pathways that promote proliferation, such
as Ceramide signaling (Gangoiti et al., 2012), G2/M transition
signaling,Cyclin signaling, and Aryl hydrocarbon receptor
signaling (Barouki et al., 2007; Yin et al., 2016), were inacti-
vated in myofiber-associated myoblasts (Figure 2B). Simi-
larly, dispersed myoblasts showed a marked inactivation
of Integrin signaling, also pointing toward inhibition of pro-
liferation (Figure 2B). Thus, canonical signaling pathways
related to cell-cycle regulation are activated or deactivated
in both culture conditions in a manner that supports prepa-
ration to exit the cell cycle. The only exception to this was
the G1/S checkpoint regulation signaling pathway, which
was activated in myofiber-associated but not dispersed myo-
blasts (Figure 2B). Similarly, the IPA category Cell Growth and
Proliferation showed a series of pro-proliferative signaling
pathways (ILK, HGF, PDGF, CREB, P70S6K, and CDC42
signaling pathways) that were activated in myofiber-associ-
ated myoblasts but deactivated in dispersed myoblasts (Fig-
ure 2C). Since proliferation and cell-cycle exit are mutually
exclusive, the fact that both pro-proliferation and pro-cell-
cycle exit pathways are activated in myofiber-associated
myoblasts, supports previous findings that myofiber-associ-
ated myoblasts are heterogeneous (Ono et al., 2010).
Lastly, we analyzed the IPA category Growth Factor
Signaling (Figure 2D). Amongst the signaling pathways that
were differentially activated between dispersed and myo-
fiber-associated myoblasts we found activation of pro-prolif-
eration signaling (HGF) and self-renewal signaling (GDNF
signaling, which requires syndecan-3 (SDC3), and therefore
is expected to promote self-renewal, Bespalov et al., 2011;
Pisconti et al., 2010) in myofiber-associated but not in
dispersed myoblasts.
Comparative Analysis of Gene Expression Suggests a
Role for the Niche in Driving Self-Renewal
To understand the differences in myoblast gene expression
changes due to culture conditions we compared ratios of
72 hr/48 hr mRNA levels within each culture condition
with each other (Figure S2C) and generated a fold-change
ratio using the formula:
Comparative ratioðCRÞ=ðmRNA at 72 hr=mRNA at 48 hrÞdispersed
ðmRNA at 72 hr=mRNA at 48 hrÞmyofiber-associated
972 Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018
Genes that score high CR values (>1) are genes that in-
crease in dispersed but not in myofiber-associated myoblasts
between 48 and 72 hr, while genes that score low CR values
(<1) are genes that increase in myofiber-associated but not in
dispersed myoblasts between 48 and 72 hr. We identified
2,583 genes with CR > 1 and 3,120 genes with CR < 1 at
1% false discovery rate (FDR), suggesting very different
time-dependent responses between the two culture condi-
tions. After filtering the resulting list of genes to include
only genes with a q value < 0.01, we then functionally
analyzed these genes by mapping them to the gene ontology
(GO) category Biological Process using the online tool
DAVID (Tables S1 and S2). Table S1 shows the GO annota-
tion of genes that had CR > 2, while Table S2 shows the
GO annotation of genes that had CR < 0.5.
The most significant GO terms to which genes that in-
crease in dispersed but not in myofiber-associated myo-
blasts (CR > 2) map are associated with cell movement
(Taxis) and muscle differentiation (Striated Muscle Differen-
tiation)(Table S1). Interestingly, the term Negative regulation
of transport also included genes that promote muscle differ-
entiation, such as Il-6 and Nos1 (De Palma et al., 2010;
Hoene et al., 2013). In contrast, the genes that increase
between 48 and 72 hr in myofiber-associated myoblasts
but not in dispersed myoblasts (CR < 0.05) map to
Cell Adhesion, Differentiation, and Cell Fate Commitment
(Table S2). The latter appears to be related mostly to main-
tenance of stemness as it contains genes that are associated
with stemness in SCs (Notch1, Notch3, Pax7, and Sox8;
Bjornson et al., 2012; Gopinath et al., 2014; Olguı
´n and
Olwin, 2004; Olguı
´n and Pisconti, 2012; Pisconti et al.,
2010; Schmidt et al., 2003).
Lastly, we organized the differentially expressed genes
according to their CR and filtered the gene list through
a manually curated list of gene families that have been
shown to play a role in myogenesis (see Experimental
Procedures for details). This analysis showed that genes
associated with muscle differentiation were increased
Figure 2. Canonical Signaling Pathways
Are Differentially Activated in Dispersed
and Myofiber-Associated Myoblasts
(A) Genes differentially expressed between
48 and 72 hr post-isolation in dispersed
(Dis) and myofiber-associated (Mf-A) myo-
blasts were functionally mapped to all ca-
nonical signaling pathways listed by IPA.
For each signaling pathway, an enrichment
p value and a Zscore of activation were
calculated and the pathways with enrich-
ment p value < 0.01 (–log[p value] > 1.3) are
plotted as heatmaps, were orange repre-
sents a positive Zscore (= activation), blue
is a negative Zscore (deactivation), and
white is Z= 0. Canonical signaling pathways
that were not enriched enough to show a
Zscore were excluded. Canonical signaling
pathways that change in opposite direction
in dispersed and myofiber-associated myo-
blasts are highlighted by a green box.
(B–D) Heatmaps obtained as in (A) for ca-
nonical signaling pathways mapping to the
IPA categories: Cell Cycle (B), Cell Growth
and Proliferation (C), and Growth Factor
Signaling (D).
(E) Heatmap distribution of the comparative
ratio (CR) = ([72 hr/48 hr]
dispersed
/[72 hr/
48 hr]
myofiber-associated
) for gene families
that are involved in myogenesis (manual
annotation, see Experimental Procedures
section).
Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018 973
in dispersed but not in myofiber-associated myoblasts
(Table S3;Figure 2E: Myog, Myl3, Myom3, Lmod3, Postn,
Ttn, and Colq). By contrast, genes associated with self-
renewal were increased in myofiber-associated but not
dispersed myoblasts (Table S3;Figure 2E: Cxcr4, Col18a1,
Notch1, Notch3, Cdc42, Pax7, Hey1, Jam2, and Sdc3).
To summarize the data shown so far: Erk1/2 downregula-
tion in dispersed myoblasts is mostly associated with acti-
vation of signaling pathways that lead to cell-cycle arrest
and the onset of differentiation. In contrast, Trp53 upregu-
lation in myofiber-associated myoblasts is associated with
a more heterogeneous transcriptomic signature in which
signaling leading to cell-cycle arrest, differentiation, self-
renewal, and proliferation co-exist. Since the role of p53
in myogenesis is complex and not well understood, we
decided to focus on understanding how an increase in
p53 levels might affect myoblast cell fate.
An Asymmetric Increase in p53 Protein Levels Is
Incompatible with Myogenic Differentiation in
Myofiber-Associated Myoblasts
The expression of p53 is rapidly and transiently upregu-
lated in the first hours of C2C12 myoblast differentiation
induced by serum deprivation (Halevy, 1993), and is impor-
tant for myogenic differentiation, as expression of a p53
dominant-negative mutant leads to impaired differentia-
tion (Soddu et al., 1996), but is not involved in differentia-
tion-induced apoptosis (Cerone et al., 2000). Thus, it could
be speculated that the upregulation of p53 that we observe
in myofiber-associated myoblasts drives differentiation.
However, it has been recently shown that a sustained in-
crease in p53 levels, such as that due to genotoxic stress,
or treatment with the MDM2 inhibitor Nutlin-3, impairs
myogenic differentiation, possibly via direct inhibition of
myogenin expression (Walsh et al., 2015; Yang et al.,
2015). Moreover, p53 mediates hypoxia-induced inhibi-
tion of myoblast differentiation (Wang et al., 2015).
To investigate the role of the increase in p53 gene expres-
sion detected by microarray in myofiber-associated myo-
blasts, we first tested whether such an increase in mRNA
levels was accompanied by an increase in p53 protein levels
and whether it was associated with differentiation. Indeed,
p53 levels appeared to increase in myofiber-associated
myoblasts, identified by the general myoblast marker
SDC3 (Cornelison et al., 2001) between 48 and 72 hr in cul-
ture, as measured by an increase in the proportion of p53+/
SDC3+ myoblasts (Figures 3A and 3B). Intriguingly, we oc-
casionally observed an asymmetric distribution of p53 in
myoblast doublets (Figure 3C), which was more prominent
at 72 hr (Figure 3D), and was accompanied by an almost
perfect inverse correlation with myogenin (MYOG) expres-
sion at 96 hr (Figure 3E). When we measured p53 protein
levels across the same 3-day time course in dispersed myo-
blasts, we found only a small and not significant increase in
p53 protein levels over time (Figure S3). Moreover, immu-
nofluorescence analysis of primary dispersed myoblasts
showed that p53 protein was still present in MYOG+
dispersed myoblasts, although often at lower levels than
in MYOG– cells (Figure 3F). These results suggest that,
when myoblasts are cultured in the presence of their native
niche, p53 protein levels increase over time dramatically
and selectively in some cells, and in these cells myogenic
differentiation is inhibited (Figure 3E). When the niche is
absent, such a dramatic and sustained increase occurs in
fewer cells, while the majority of cells maintain lower levels
of p53, which are permissive of myogenic differentiation
(Figure 3F).
If it is true that the presence of the SC niche promotes
p53 upregulation in a subpopulation of myofiber-associ-
ated myoblasts and that this in turn inhibits myogenic dif-
ferentiation in these cells, then decreasing p53 in myofiber-
associated myoblasts via RNAi should promote myoblast
differentiation in myofiber cultures. To test this hypothe-
sis, we transfected myofiber cultures with either p53 small
interfering RNA (siRNA) or with control (scrambled) siRNA
at 48 hr post-isolation. Transfection of C2C12 myoblasts
was used to validate p53 knockdown efficiency (Figure S4).
Two days after transfection, we fixed and immunostained
the myofibers to detect and quantify MYOG+ myoblasts
(Figure 4A). As predicted, the percentage of MYOG+/
SDC3+ cells was increased in p53 siRNA-transfected cul-
tures compared with control siRNA-transfected cultures,
although the difference did not reach statistical signifi-
cance (Figure 4B). The fact that the difference was small,
could be due to the high propensity of differentiated myo-
fiber-associated myoblasts to immediately fuse with the
underlying fiber (Kuang et al., 2007), which removes
them from the equation when MYOG+/SDC3+ cells are
scored as a measure of differentiation. Indeed, the number
of myonuclei per unit length was significantly increased
(Figure 4C), while the frequency of myoblasts (SDC3+ cells)
per unit length was decreased (Figure 4D) upon p53 knock-
down, supporting the idea that p53 inhibits differentiation
in a subpopulation of myofiber-associated myoblasts.
The next question was: what is the fate of these myo-
blasts that show high p53 protein levels and fail to differen-
tiate? Since high levels of p53 are incompatible with prolif-
eration, we hypothesized that a sustained increase in p53
levels in myoblasts would lead to apoptosis, senescence,
or quiescence.
An Increase in p53 Protein Levels Promotes Myoblast
Quiescence
Nutlin-3, a small molecule compound that blocks the inter-
action between p53 and the E3 ligase MDM2, is often used
to increase the levels of p53 protein in the cell (Vassilev,
974 Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018
2005). Nutlin-3 treatment of C2C12 myoblasts leads to a
generalized increase in p53 protein levels and both prolifer-
ation and differentiation inhibition (Walsh et al., 2015),
thus supporting our hypothesis that a sustained increase
in p53 levels in primary myoblasts might lead to apoptosis,
senescence, or quiescence, as well as differentiation inhibi-
tion. To test this hypothesis we treated primary dispersed
myoblasts with increasing concentrations of Nutlin-3 and
verified that indeed Nutlin-3 inhibits proliferation (Fig-
ure S5A). However, Nutlin-3 treatment does not seem to in-
crease apoptosis in primary myoblasts (Figure S5B), which
is consistent with the absence of a role for p53 in serum
deprivation-induced myoblast apoptosis (Cerone et al.,
2000). Since Nutlin-3 does not promote differentiation,
in fact it inhibits differentiation (Figure S6), the lack of
apoptosis suggests that Nutlin-3-mediated inhibition of
Figure 3. p53 Increases over Time in a
Subset of Myofiber-Associated Myoblasts
and Is Asymmetrically Distributed
(A) Individual myofibers were isolated and
cultured in suspension for 48 and 72 hr prior
to fixation and immunostaining to detect
p53 (green), the myoblast marker synde-
can-3 (SDC3, red), and DNA (DAPI, blue).
Arrows indicate SDC3+ SCs.
(B) Quantification of (A) where at least
15 myofibers across 3 independent experi-
ments had been scored, and the percentage
of p53+ cells over the total number of SDC3+
cells plotted as average ±SEM. **p < 0.01.
(C) Individual myofibers were isolated and
cultured as in (A) then immunostained
to detect p53 (green), SDC3 (white), and
DNA (PI, red). Arrows indicate doublets of
dividing cells where p53 is distributed
either symmetrically (48 hr, left panels) or
asymmetrically (72 hr, right panels).
(D) Quantification of (C) where at least 15
fibers across 3 independent experiments
had been scored.
(E and F) p53+/MYOG– myoblasts are very
abundant in myofiber cultures (E) but less in
dispersed cultures (F). Individual myofibers
or primary myoblasts were isolated and
cultured for 96 hr, then fixed and immuno-
stained to detect p53 (green), myogenin
(MYOG, red), and DNA (DAPI, blue). Arrows
in (E) indicate p53+/MYOG+ cells, while
all the other cells are either p53+/MYOG–
or p53-/MYOG+. Arrowheads in (F) indicate
p53+/MYOG– cells. All the other cells are
p53+/MYOG+.
Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018 975
proliferation is due to induction of either quiescence or
senescence.
To discriminate between quiescence and senescence we
performed a reserve cell assay. Cultures of both primary
dispersed myoblasts and C2C12 myoblasts provide a
validated ex vivo model of SC self-renewal: upon serum
lowering, the vast majority of myoblasts exit the cell cycle
with a portion of them downregulating PAX7 and undergo-
ing terminal differentiation, and another portion maintain-
ing high levels of PAX7 and becoming ‘‘reserve cells’ (Yosh-
ida et al., 1998). In reserve cell assays (Figure 5A), Nutlin-3
treatment of both primary and C2C12 myoblast cultures
dramatically increased the percentage of putative reserve
cells (PAX7+/KI67–) generated by serum lowering (Figures
5B, 5C, 5E, and 5F). This was accompanied by a decrease
in the percentage of proliferating myoblasts (PAX7+/
KI67+) in C2C12 cultures (Figures 5E and 5G), although
no significant changes in PAX7+/KI67+ cells were detected
in primary myoblast cultures (Figures 5B and 5D).
In addition to being mitotically quiescent, reserve cells
also show decreased levels of MYOD (Yoshida et al.,
1998), thus ‘‘true’ reserve cells are PAX7+/KI67–/MYOD–
(Yue et al., 2016). Although we could effectively label
C2C12 myoblasts to simultaneously detect PAX7, KI67,
and MYOD, this could not be achieved with primary myo-
blasts (triple PAX7/KI67/MYOD staining of primary myo-
blasts showed cytoplasmic positivity for MYOD and/or
KI67, even at very low antibody concentration, which is
not observed when either KI67 or MYOD antibodies are
used in a double staining of C2C12 cells with PAX7).
Thus, we were forced to label primary myoblasts separately
for PAX7 and KI67 (as discussed above) and PAX7 and
MYOD. Nutlin-3 treatment dramatically reduced MYOD
protein levels and led to an increase in the frequency
of true reserve cells (PAX7+/KI67–/MYOD–) in C2C12
myoblast cultures (Figures 5K–5L), which was accompa-
nied by a decrease in proliferating myoblasts (PAX7+/
MYOD+/KI67+; Figures 5K and 5M). Similarly, Nutlin-3
treatment produced a 20-fold increase in the numbers of
uncommitted myoblasts (PAX7+/MYOD– cells; Figures
5H and 5I), but only a 6-fold increase in the percentage of
committed myoblasts (PAX7+/MYOD+; Figures 5H and
5J) in primary myoblast cultures. Although in the absence
of a triple PAX7/MYOD/KI67 staining it is not possible to
establish whether Nutlin-3 promotes true reserve cell gen-
eration in primary myoblast cultures as it does in C2C12
cultures, the observations that: (1) the increase in uncom-
mitted myoblasts produced by Nutlin-3 is over 3 times
greater than the increase in committed myoblasts (20-
versus 6-fold), and (2) PAX7+/KI67+ myoblasts do not
increase in response to Nutlin-3 in primary cultures (Fig-
ure 5D), while PAX7+/KI67– cells increase 7-fold (Fig-
ure 5C), suggest that the percentage of true reserve cells
(PAX7+/KI67–/MYOD–) is likely to significantly increase
in primary myoblast cultures as it does in C2C12 cultures
in response to Nutlin-3 treatment.
In contrast to differentiated and senescent cells, which
are in an irreversible G0 phase, reserve cells are quiescent,
Figure 4. p53 Knockdown Promotes Myoblast Differentiation in Myofiber Cultures
(A) Individual myofibers were isolated and cultured in suspension for 48 hr prior to transfection with either a specific p53 siRNA or
a scrambled (ctrl) siRNA and 4 hr later fixed and immunostained to detect myogenin (MYOG, green), syndecan-3 (SDC3, red), and DNA
(DAPI, blue).
(B) Differentiated myoblasts were scored as percentage of MyoG+/SDC3+ myoblasts over total nuclei per unit length across at least eight
myofibers/experiment in three independent experiments (N > 24), and plotted as average ±SEM.
(C) The number of myonuclei per unit length was measured as total number of DAPI+ nuclei minus the number of nuclei contained in SDC3+
cells (since SDC3 marks SCs and myoblasts at all stages in myogenesis) per unit length across at least eight myofibers/experiment in three
independent experiments (N > 24), and plotted as average ±SEM.
(D) The percentage of myoblasts was calculated as percentage of nuclei contained in SDC3+ cells over the total number of DAPI+ nuclei per
unit length across at least eight myofibers/experiment in three independent experiments (N > 24), and plotted as average ±SEM.
**p < 0.01 when comparing the indicated population scored in control siRNA-transfected cultures with the same population scored in p53
siRNA-transfected cultures.
976 Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018
Figure 5. Nutlin-3 Promotes Reserve Cell Generation in Serum-Deprived Myoblasts
(A) Schematic representation of the reserve cell assay experimental design. Primary and C2C12 myoblast cultures were switched to low
serum medium to induce cell-cycle exit in the presence of either 20 mM Nutlin-3 (Nut3) or vehicle (DMSO). To distinguish between
quiescence and senescence, myoblast cultures that had been maintained in low serum supplemented with either DMSO or Nutlin-3 for
3 days were re-exposed to high serum for 2 days. The experiment was repeated 3 times independently and each time 10–15 technical
replicates were scored.
(B–D) Primary myoblast cultures were treated as in (A) and, after 3 days in low serum, fixed, immunostained to detect PAX7 (green), KI67
(red), and DNA (DAPI, blue), and scored for the percentages of PAX7+/KI67– (C) and PAX7+/KI67+ (D) cells. In (B), one representative
image for each treatment is shown. In (C) and (D) quantitative analyses of the indicated cell subpopulations across all three independent
experiments were plotted as average ±SEM. The arrows in (B) indicate PAX7+/KI67+ cells.
(E–G) C2C12 myoblasts cultures were treated as in (A) and, after 3 days in low serum, fixed, and immunostained as in (B). In (E) one
representative image for each treatment is shown. In (F) and (G) quantitative analyses of the indicated cell subpopulations across all three
independent experiments were plotted as average ±SEM.
(H–M) Primary (H and J) and C2C12 (K–M) myoblasts that had been maintained in low serum supplemented with either DMSO or Nutlin-3
for 3 days were fixed and immunostained to detect PAX7 (green), KI67, and/or MYOD (red) and DNA (DAPI, blue). In (H) and (K) one
representative image for each treatment is shown. Insets are enlarged on the side of the main image to show mutual exclusion or
co-localization of PAX7 and MYOD (H) or PAX7 and MYOD+ KI67 (K). In (I), (J), (L), and (M) quantitative analyses of the indicated cell
subpopulations across all three independent experiments were plotted as average ±SEM.
(N–Q) In order to quantify true reserve cells, after 3 days in low serum, primary (N and O) and C2C12 (P and Q) myoblast cultures were
re-exposed to high serum as depicted in (A) for 2 days, then fixed and immunostained to detect PAX7, KI67, and DNA. The percentages of
PAX7+/KI67– (N and P) and PAX7+/KI67+ (O and Q) were calculated and plotted as fold change of the same population in cultures that had
been maintained in low serum and then re-exposed to high serum (we will call it ‘post-wash’’ for simplicity) versus the same population in
cultures maintained for only 3 days in low serum (indicated as ‘‘pre-wash’’). The average of the post-wash/pre-wash fold change for each
(legend continued on next page)
Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018 977
in a reversible G0 phase, and therefore can be induced to re-
enter the cell cycle by adding serum to the culture medium
(Yoshida et al., 1998). To discriminate whether the putative
reserve cells generated in myoblast cultures in response to
Nutlin-3 treatment were truly quiescent or senescent, we
exposed again to high serum the cultures that had been
maintained for 3 days in low serum supplemented with
either Nutlin-3 or DMSO, and 2 days later we measured
the percentages of cells that had re-entered the cell cycle
(Figure 5A). In both C2C12 and primary myoblast cultures
washed and re-exposed to high serum, the percentage of
PAX7+/KI67+ cells increased more dramatically if the cul-
tures had been previously treated with Nutlin-3 compared
with vehicle (Figures 5O and 5Q). Interestingly, the per-
centage of PAX7+/KI67– cells did not significantly change
upon serum re-exposure in cultures that had been previ-
ously treated with vehicle, but significantly decreased
in cell cultures that had been previously treated with
Nutlin-3 (Figures 5N and 5P). These results indicate that
the percentage of true reserve cells is normally very low in
our reserve cell assay, regardless of the origin of the myo-
blasts used (primary or C2C12). However, this percentage
of true reserve cells is dramatically increased by Nutlin-3
treatment.
An Increase in p53 Levels Is Directly Responsible for
the Increased Reserve Cell Generation in Response to
Nutlin-3
To test whether Nutlin-3 promoted reserve cell generation
and inhibited differentiation directly via p53, we trans-
fected C2C12 myoblasts with either a control siRNA or
with an anti-p53 siRNA, and 3 hr later we replaced the
transfection medium with low serum medium (to induce
cell-cycle exit followed by myotube and reserve cell
generation) supplemented with either Nutlin-3 or vehicle.
Although p53 siRNA transfection did not appear to alter
the differentiation capacity of C2C12 myoblasts treated
with vehicle, it partly rescued the differentiation levels of
cells treated with Nutlin-3 (Figure 6A). More importantly,
the percentage of PAX7+/KI67– reserve cells that was
increased by Nutlin-3 in cultures transfected with control
siRNA, was then rescued by transfection with a specific
anti-p53 siRNA (Figures 6B and 6C), further indicating
a direct role for p53 in regulating the balance between dif-
ferentiation and reserve cell generation. Interestingly, the
decrease in the percentage of PAX7+/KI67– reserve cells
caused by p53 knockdown in Nutiln-3-treated cultures
was not due to a rescue of the number of proliferating
(PAX7+/KI67+) myoblasts (Figure 6B), but to a decrease
in differentiation. This suggests that Nutlin-3 promotes
reserve cell generation via p53 by affecting directly the
cell fate decision to either differentiate or become quies-
cent, rather than the earlier decision to either divide or
exit the cell cycle.
DISCUSSION
In this study, we show that primary myoblasts cultured
ex vivo carry out different transcriptional programs to regu-
late their cell fate transitions according to whether they
are cultured in the presence or absence of their native
niche. In the absence of the niche, the vast majority of
myoblasts activate a transcriptional program dominated
by Erk1/2 downregulation and differentiate. In the pres-
ence of the niche, the transcriptional program is domi-
nated by an upregulation of Trp53. Validation experiments
revealed that a sustained p53 increase in myofiber-associ-
ated myoblasts is restricted to a subpopulation of cells
that fail to differentiate. These myoblasts showing high
levels of p53 are more rare in the absence of the niche,
which is likely why p53 upregulation is not significantly
detected in the transcriptomic analysis of the dispersed
myoblasts. Thus, our data suggest that the presence of
the niche leads to a greater percentage of myoblasts
showing a sustained increase in p53 levels and associated
fail to differentiate. We then used gain- and loss-of-
function experiments in a culture context that promotes
cell-cycle exit to determine the fate of these high p53
myoblasts that fail to differentiate, and showed that for
the most part these myoblasts become quiescent. Our
data therefore point toward p53 signaling as a regulator
of the balance between differentiation and self-renewal:
although a transient increase in p53 levels occurs in the
early hours that precede differentiation and is likely neces-
sary for cell-cycle exit, p53 must quickly return to basal
levels upon cell-cycle exit in order for differentiation to
occur (Figure 7). However, in some cells, p53 levels further
increase and remain high over time, leading to differenti-
ation inhibition and promotion of self-renewal (Figure 7).
The latter process appears to be promoted when myogen-
esis occurs in the presence of the SC niche.
subpopulation across 3 independent biological replicates (each one scored for 10–15 technical replicates) was calculated and
plotted ±SEM.
In (C) and (D), (F) and (G), (I) and (J), and (L) and (M): **p < 0.01, where p is the p value of the average percentage of each cell
subpopulation in Nut-3-treated cultures versus DMSO-treated cultures. In (N)–(Q),
#
p < 0.05 and
##
p < 0.01, where p is the p value of the
fold change calculated as ‘‘subpopulation percentage in post-wash/subpopulation percentage in pre-wash’ within each treatment (DMSO
or Nut-3).
978 Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018
Our findings open up new questions regarding the mo-
lecular mechanisms that underlie p53-mediated regulation
of the balance between differentiation and self-renewal.
One possibility is that p53 regulates such balance via direct
inhibition of myogenin expression (Yang et al., 2015) and
consequently inhibition of differentiation, upregulation
of PAX7, and activation of a program that resembles in vivo
quiescence (Olguı
´n and Olwin, 2004). By contrast, a direct
upregulation of PAX7 by p53 appears unlikely (Yang et al.,
2015). Interestingly, mice lacking the p53 target gene PW1
are born with fewer quiescent SCs (Nicolas et al., 2005).
Since it is established that neonatal SCs are derived during
development form MYOD+ myoblasts (Kanisicak et al.,
2009), the finding that PW1 mutants lack quiescent SCs
at birth further supports our view that p53 promotes gener-
ation of a population of quiescent SCs from proliferating
myoblasts.
Although it had been previously shown that p53 inacti-
vation via expression of a dominant-negative mutant de-
creases myoblast differentiation (Soddu et al., 1996) and
that differentiation is decreased in primary Trp53
/
myo-
blasts (Porrello et al., 2000), when we decreased p53 levels
in myoblasts by transfecting a specific anti-p53 siRNA we
did not observe altered differentiation in the absence of
Nutlin-3. This might be due to the fact that we lowered
serum levels prior to an effective knock down of p53 (the
medium was switched only 3 hr after transfection). Indeed,
also reserve cell generation was not affected by p53 knock-
down in cultures that had not been treated with Nutlin-3.
Thus, it is possible that a transient peak of p53 expression
still occurred in cultures that had not been treated with
Nutlin-3 prior to effective p53 knockdown by siRNA,
thus ensuring a normal distribution between differentia-
tion and quiescence. Vice versa, in cultures treated with
Nutlin-3, p53 levels either rapidly increased in response
to Nutlin-3 above the normal transient peak and remained
high for the entire duration of the experiment (control
siRNA), or, in the absence of Nutlin-3, were prevented
from increasing further than the normal transient peak
by p53 siRNA, and in this second case the distribution
between differentiation and quiescence was maintained
at basal levels. Thus, our data suggest that, after cell-cycle
exit, the levels of p53 in the cell determine whether the
cell is going to enter a terminal G0 phase and differentiate
(p53 levels returning low; Figure 7), or whether the cell is
going to remain in a reversible G0 phase and effectively
renew the pool of undifferentiated, quiescent SCs (p53
levels further increasing and remaining high; Figure 7).
Overall our study supports a model whereby myogenic
differentiation is the preferred fate choice once SCs have
been removed from their anatomical niche and cultured
as myoblasts, whereas the presence of the niche provides
signals that promote quiescence and self-renewal via mul-
tiple mechanisms, one of which leads to p53 upregulation.
Further research is necessary in order to identify the molec-
ular mechanisms through which the SC niche regulates
Figure 6. p53 Knockdown Rescues
Reduced Differentiation and Increased
Reserve Cell Generation in Myoblasts
Treated with Nutlin-3
(A) Differentiation of C2C12 myoblasts is
reduced by Nutlin-3 treatment and rescued
by p53 knockdown. C2C12 myoblasts were
transfected with either control or p53
siRNA, then maintained for 3 days in low
serum supplemented with either DMSO
or 20 mM Nutlin-3 prior to fixation and
immunostaining to detect myosin heavy
chain (MYHC, green) and DNA (DAPI,
blue).
(B) Reserve cell (PAX7+/KI67–) generation
is increased by Nutlin-3 treatment and
rescued by p53 knockdown. C2C12 myo-
blasts were transfected and cultured as in
(A) prior to fixation and immunostaining to
detect PAX7 (green), KI67 (red), and DNA
(DAPI, blue). Arrows point to PAX7+/KI67+
cells. All PAX7+ cells not labeled by an
arrow are KI67–. All KI67+ cells that are not
labeled by an arrow are PAX7–.
(C) Quantification of (B) where the average percentage of PAX7+/KI67– cells (over total DAPI+ cells) is calculated across 10–15 technical
replicates for 3 independent experiments (N = 35) and plotted. Error bars are SEM. **p < 0.01.
Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018 979
p53 protein levels in myoblasts; however, our results
strongly support the notion that p53 plays an important
physiological role in muscle stem cell biology.
EXPERIMENTAL PROCEDURES
Mice
Male C57Bl/6J mice were purchased from Charles River and
housed until used (at 12–13 weeks of age) in a pathogen-free facil-
ity at the University of Liverpool, in accordance with the Animals
(Scientific Procedures) Act 1986 and the EU Directive 2010/63/EU,
and after local ethical review and approval by the Animal Welfare
and Ethical Review Body.
Myofiber Cultures
Single myofibers were isolated and cultured as described previously
(Pisconti et al., 2016). Further details can be found in Supplemental
Experimental Procedures.
Primary Dispersed Myoblast Cultures
Dispersed myoblasts were prepared and cultured as described
previously (Pisconti et al., 2010). Further details can be found in
Supplemental Experimental Procedures.
C2C12 Cell Cultures
C2C12 myoblasts (Yaffe and Saxel, 1977) were cultured as
described previously (Arecco et al., 2016). Further details can be
found in Supplemental Experimental Procedures.
Immunofluorescence
Myofibers, dispersed primary, and C2C12 myoblasts were fixed
for either 20 min (myofibers) or 10 min (primary and C2C12
myoblasts) at room temperature with 4% paraformaldehyde prior
to being processed for immunofluorescence as described previ-
ously (Pisconti et al., 2016). For a list of primary and secondary
antibodies, the concentrations used and microscopy details, refer
to Supplemental Experimental Procedures.
Isolation of Myoblasts from Myofiber Cultures
Forty-eight and 72 hr after isolation, myofibers were collected,
washed twice with PBS, always collecting the supernatant which
contained loosely attached myoblasts, then treated with 0.05%
trypsin for 10 min at 37C. The action of trypsin was then
stopped by addition of 2 volumes of primary SC growth medium
and the myofibers separated from stripped myoblasts by centrifu-
gation (3 min at 100 3g): the supernatant was collected and
combined with the two previous PBS washes. Myoblasts were
then collected by centrifugation at 500 3g, the pellet washed
once in PBS, and then immediately lysed in RLT buffer (RNeasy
Kit, QIAGEN). This method yielded highly pure preparations of
myofiber-associated myoblasts (Figures S6A–S6E) and high-qual-
ity RNA (Figures S6F and S6G), which did not contain contami-
nant RNA from the myofiber (see Supplemental Information for
details).
RNA Extraction and Quality Control
RNA was extracted using the RNeasy Kit (QIAGEN) according
to the manufacturer’s instructions. The quality and concentration
of the RNA was assessed using a Bioanalyzer (Agilent 2100, Agilent)
or a Nanodrop (Thermo Fisher).
Microarrays
Sample labeling and microarray hybridizations were carried
out according to Agilents One-Colour Microarray-Based Gene
Expression Analysis Protocol, v.6.6 (manual part number
G4140-90040). Further details can be found in the Supplemental
Experimental Procedures. Raw microarray data are available from
Figure 7. Schematic Representation of
the Role of p53 in SC-Mediated Myo-
genesis
In uninjured muscle (red myofiber on the
left), SCs (oval yellow cell on the my-
ofiber) are quiescent, in a reversible G0
phase of the cell cycle. Upon injury (bolt)
SCs become activated and re-enter the
cell cycle in G1 (yellow/green, star-shaped
cell). At each division cycle the two
daughter cells (myoblasts, green rounded
cells) choose between three main fates:
divide again (and complete the cell cycle),
or exit the cell cycle and then either
differentiate (red elongated cells), or
become quiescent (again oval yellow cell).
Although a transient increase in p53 levels
(upward blue arrow) allows cell-cycle exit,
p53 must return to basal levels in order
for differentiation to proceed (downward
blue arrow). In contrast, a further and sustained increase in p53 levels (larger upward blue arrow) leads to differentiation
inhibition and promotion of self-renewal (return to a quiescent, undifferentiated state).
980 Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018
the GEO public depository under the accession number: GEO:
GSE109052.
Microarray Analysis
Raw Agilent text files were read using the read. Agilent function in
the marray R package (Yang, Paquet, and Dudoit, Exploratory anal-
ysis for two-color spotted microarray data. R package v.1.50.0.
http://www.maths.usyd.edu.au/u/jeany/). Data were background
corrected and loess normalized (using the normalizeWithinArrays
function within the Limma package, followed by a quantile
normalization using the normalizeBetweenArrays function)
against their respective 48-hr data point to allow for a direct
comparison between the two culture systems at both time points.
Data were then summarized at the gene level by first averaging
identical probes and then selecting a representative probe per
gene name that showed the highest average intensity across the
dataset. Genes with a median intensity level <5 across all arrays
were removed. Differentially expressed genes were identified
using the samr package (Tusher et al., 2001) for each of the three
comparisons: (1) 72 versus 48 hr myofiber-associated myoblasts
(Figure S2A), (2) 72 versus 48 hr dispersed myoblasts (Figure S2B),
and (3) 72 hr/48 hr dispersed versus 72 hr/48 hr myofiber-associ-
ated myoblasts (Figure S2C). Genes significant at 1% FDR in each
comparison were extracted and used for further analysis.
Manual Annotation of Gene Families Involved in
Myogenesis
Gene families involved in myogenesis were identified manually via
screening of literature published in PubMed between 01/01/1986
and 01/01/2016. A gene family was annotated as being involved
in myogenesis when at least one member of the family was
described as involved in myogenesis in at least one peer-reviewed
publication. This list included but was not limited to: myogenesis
regulatory factors, Pax genes, skeletal muscle sarcomeric proteins,
various extracellular matrix components, calcitonin receptor,
Wnt pathway, Notch pathway, receptor tyrosine kinase pathways
(fibroblast growth factor, hepathocyte growth factor, epidermal
growth factor, insulin growth factor, etc.), transforming growth
factor bfamily pathways, interleukins, cytokines, chemokines,
mitogen-activated protein kinases, p53 family, cyclins, cyclin-
dependent kinases, etc. The comparative analysis filtered through
this manually curated list is reported in full in Table S3.
Transfection
Myofiber and C2C12 cell cultures were transfected as described
previously (Ghadiali et al., 2016). Further details can be found in
Supplemental Experimental Procedures.
Western Blotting
Primary and C2C12 myoblasts were lysed in modified RIPA buffer
(50 mM Tris-HCl [pH 7.5], 150 mM NaCl, 1% IGEPAL, 0.5% SDS,
and 0.1% sodium deoxycholate) and western blotting performed
as described previously (Arecco et al., 2016). A list of primary and
secondary antibodies and concentrations used can be found in
the Supplemental Experimental Procedures. Quantitative analysis
of western blots was performed using the ‘‘Analyze Gel’ function
of ImageJ.
Statistical Analysis of Cell and Biochemistry
Experiments
Experiments shown in Figures 3,4,5, and 6were statistically
analyzed using a t test as the data distributed normally. A p value <
0.05 was considered significant.
SUPPLEMENTAL INFORMATION
Supplemental Information includes Supplemental Experimental
Procedures, seven figures, and three tables and can be found
with this article online at https://doi.org/10.1016/j.stemcr.2018.
01.007.
AUTHOR CONTRIBUTIONS
V.F. designed and performed experiments shown in Figures 1,2,
and S2. R.S.G. designed and performed experiments shown in Fig-
ures 4 and S3. P.A. performed the microarray analysis and the CR
analysis. A.R. performed experiments shown in Figure 5. J.E.T.
contributed to experimental design. A.P. contributed to experi-
mental design and execution throughout the study and drafted
the manuscript. All authors contributed to manuscript preparation
by providing critical feedback.
ACKNOWLEDGMENTS
This work was funded by a Marie Curie IEF (grant no. 302113) and
a Wellcome Trust ISSF to A.P., a BBSRC-DTP PhD studentship to
R.S.G., and Natural Environment Research Council grant to P.A.
(NE/M01939X/1). We wish to thank Prof. Dave Fernig for the
FGF2, Prof. Brad Olwin for the anti-SDC3 antibody, and Ms. Lisa
Olohan and Ms. Pia Koldkjaer, University of Liverpool, for their
help with the microarrays. We also wish to thank the staff of the
Biomedical Services Unit at the University of Liverpool for the su-
perb care they take of the mouse colonies.
Received: October 11, 2016
Revised: January 11, 2018
Accepted: January 12, 2018
Published: February 8, 2018
REFERENCES
Arecco, N., Clarke, C.J., Jones, F.K., Simpson, D.M., Mason, D.,
Beynon, R.J., and Pisconti, A. (2016). Elastase levels and activity
are increased in dystrophic muscle and impair myoblast cell sur-
vival, proliferation and differentiation. Sci. Rep. 6, 24708.
Barouki, R., Coumoul, X., and Fernandez-Salguero, P.M. (2007).
The aryl hydrocarbon receptor, more than a xenobiotic-interacting
protein. FEBS Lett. 581, 3608–3615.
Bespalov, M.M., Sidorova, Y.A., Tumova, S., Ahonen-Bishopp, A.,
Magalha
˜es, A.C., Kulesskiy, E., Paveliev, M., Rivera, C., Rauvala,
H., and Saarma, M. (2011). Heparan sulfate proteoglycan synde-
can-3 is a novel receptor for GDNF, neurturin, and artemin.
J. Cell Biol. 192, 153–169.
Bjornson, C.R.R., Cheung, T.H., Liu, L., Tripathi, P.V., Steeper, K.M.,
and Rando, T.A. (2012). Notch signaling is necessary to maintain
quiescence in adult muscle stem cells. Stem Cells 30, 232–242.
Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018 981
Cerone, M.A., Marchetti, A., Bossi, G., Blandino, G., Sacchi, A., and
Soddu, S. (2000). p53 is involved in the differentiation but not in
the differentiation-associated apoptosis of myoblasts. Cell Death
Differ. 7, 506–508.
Cornelison, D.D., Filla, M.S., Stanley, H.M., Rapraeger, A.C., and
Olwin, B.B. (2001). Syndecan-3 and syndecan-4 specifically mark
skeletal muscle satellite cells and are implicated in satellite cell
maintenance and muscle regeneration. Dev. Biol. 239, 79–94.
Dumont, N.A., Wang, Y.X., and Rudnicki, M.A. (2015). Intrinsic
and extrinsic mechanisms regulating satellite cell function. Devel-
opment 142, 1572–1581.
Gangoiti, P., Bernacchioni, C., Donati, C., Cencetti, F., Ouro, A.,
Go
´mez-Mun
˜oz, A., and Bruni, P. (2012). Ceramide 1-phosphate
stimulates proliferation of C2C12 myoblasts. Biochimie 94,
597–607.
Ghadiali, R.S., Guimond, S.E., Turnbull, J.E., and Pisconti, A.
(2016). Dynamic changes in heparan sulfate during muscle differ-
entiation and ageing regulate myoblast cell fate and FGF2 signal-
ling. Matrix Biol. 59, 54–68.
Gopinath, S.D., Webb, A.E., Brunet, A., and Rando, T.A. (2014).
FOXO3 promotes quiescence in adult muscle stem cells during
the process of self-renewal. Stem Cell Reports 2, 414–426.
Halevy, O. (1993). p53 gene is up-regulated during skeletal
muscle cell differentiation. Biochem. Biophys. Res. Commun.
192, 714–719.
Hoene, M., Runge, H., Ha
¨ring, H.U., Schleicher, E.D., and Weigert,
C. (2013). Interleukin-6 promotes myogenic differentiation of
mouse skeletal muscle cells: role of the STAT3 pathway. Am. J.
Physiol. Cell Physiol. 304, C128–C136.
Jones, N.C., Fedorov, Y.V., Rosenthal, R.S., and Olwin, B.B. (2001).
ERK1/2 is required for myoblast proliferation but is dispensable
for muscle gene expression and cell fusion. J. Cell Physiol. 186,
104–115.
Kanisicak, O., Mendez, J.J., Yamamoto, S., Yamamoto, M., and
Goldhamer, D.J. (2009). Progenitors of skeletal muscle satellite
cells express the muscle determination gene, MyoD. Dev. Biol.
332, 131–141.
Kuang, S., Kuroda, K., Le Grand, F., and Rudnicki, M.A. (2007).
Asymmetric self-renewal and commitment of satellite stem cells
in muscle. Cell 129, 999–1010.
Lazaro, J.B., Kitzmann, M., Poul, M.A., Vandromme, M., Lamb,
N.J., and Fernandez, A. (1997). Cyclin dependent kinase 5, cdk5,
is a positive regulator of myogenesis in mouse C2 cells. J. Cell
Sci. 110 (Pt 10), 1251–1260.
Levine, A.J. (1997). p53, the cellular gatekeeper for growth and di-
vision. Cell 88, 323–331.
Mashinchian, O., Pisconti, A., Le Moal, E., and Bentzinger, F.C.
(2017). The muscle stem cell niche in health and disease. Curr.
Top. Dev. Biol. Published online November 24, 2017. https://doi.
org/10.1016/bs.ctdb.2017.08.003.
Nicolas, N., Marazzi, G., Kelley, K., and Sassoon, D. (2005). Embry-
onic deregulation of muscle stress signaling pathways leads to
altered postnatal stem cell behavior and a failure in postnatal mus-
cle growth. Dev. Biol. 281, 171–183.
Olguı
´n, H.C., and Olwin, B.B. (2004). Pax-7 up-regulation inhibits
myogenesis and cell cycle progression in satellite cells: a potential
mechanism for self-renewal. Dev. Biol. 275, 375–388.
Olguı
´n, H.C., and Pisconti, A. (2012). Marking the tempo for myo-
genesis: Pax7 and the regulation of muscle stem cell fate decisions.
J. Cell. Mol. Med. 16, 1013–1025.
Ono, Y., Boldrin, L., Knopp, P., Morgan, J.E., and Zammit, P.S.
(2010). Muscle satellite cells are a functionally heterogeneous pop-
ulation in both somite-derived and branchiomeric muscles. Dev.
Biol. 337, 29–41.
De Palma, C., Falcone, S., Pisoni, S., Cipolat, S., Panzeri, C., Pam-
bianco, S., Pisconti, A., Allevi, R., Bassi, M.T., Cossu, G., et al.
(2010). Nitric oxide inhibition of Drp1-mediated mitochondrial
fission is critical for myogenic differentiation. Cell Death Differ.
17, 1684–1696.
Pisconti, A., Cornelison, D.D.W., Olguı
´n, H.C., Antwine, T.L., and
Olwin, B.B. (2010). Syndecan-3 and Notch cooperate in regulating
adult myogenesis. J. Cell Biol. 190, 427–441.
Pisconti, A., Banks, G.B., Babaeijandaghi, F., Betta, N.D., Rossi,
F.M.V., Chamberlain, J.S., and Olwin, B.B. (2016). Loss of niche-
satellite cell interactions in syndecan-3 null mice alters muscle
progenitor cell homeostasis improving muscle regeneration.
Skelet. Muscle 6,34.
Porrello, A., Cerone, M.A., Coen, S., Gurtner, A., Fontemaggi, G.,
Cimino, L., Piaggio, G., Sacchi, A., and Soddu, S. (2000). p53 regu-
lates myogenesis by triggering the differentiation activity of pRb.
J. Cell Biol. 151, 1295–1304.
Sarker, K.P., and Lee, K.Y. (2004). L6 myoblast differentiation is
modulated by Cdk5 via the PI3K-AKT-p70S6K signaling pathway.
Oncogene 23, 6064–6070.
Schmidt, K., Glaser, G., Wernig, A., Wegner, M., and Rosorius, O.
(2003). Sox8 is a specific marker for muscle satellite cells and in-
hibits myogenesis. J. Biol. Chem. 278, 29769–29775.
Soddu, S., Blandino, G., Scardigli, R., Coen, S., Marchetti, A., Rizzo,
M.G., Bossi, G., Cimino, L., Crescenzi, M., and Sacchi, A. (1996).
Interference with p53 protein inhibits hematopoietic and muscle
differentiation. J. Cell Biol. 134, 193–204.
Troy, A., Cadwallader, A., Fedorov, Y., Tyner, K., Tanaka, K., and
Olwin, B. (2012). Coordination of satellite cell activation and
self-renewal by par complex-dependent asymmetric activation of
p38alpha/beta MAPK. Cell Stem Cell 11, 541–553.
Tusher, V.G., Tibshirani, R., and Chu, G. (2001). Significance anal-
ysis of microarrays applied to the ionizing radiation response. Proc.
Natl. Acad. Sci. USA 98, 5116–5121.
Vassilev, L.T. (2005). p53 Activation by small molecules: applica-
tion in oncology. J. Med. Chem. 48, 4491–4499.
Walsh, E.M., Niu, M., Bergholz, J., and Xiao, Z.-X.J. (2015).
Nutlin-3 down-regulates retinoblastoma protein expression and
inhibits muscle cell differentiation. Biochem. Biophys. Res. Com-
mun. 461, 293–299.
Wang, C., Liu, W., Liu, Z., Chen, L., Liu, X., and Kuang, S. (2015).
Hypoxia inhibits myogenic differentiation through p53 protein-
dependent induction of Bhlhe40 protein. J. Biol. Chem. 290,
29707–29716.
982 Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018
Wen, Y., Bi, P., Liu, W., Asakura, A., Keller, C., and Kuang, S. (2012).
Constitutive Notch activation upregulates Pax7 and promotes the
self-renewal of skeletal muscle satellite cells. Mol. Cell Biol. 32,
2300–2311.
Yaffe, D., and Saxel, O. (1977). Serial passaging and differentiation
of myogenic cells isolated from dystrophic mouse muscle. Nature
270, 725–727.
Yang, Z.J.P., Broz, D.K., Noderer, W.L., Ferreira, J.P., Overton, K.W.,
Spencer, S.L., Meyer, T., Tapscott, S.J., Attardi, L.D., and Wang, C.L.
(2015). p53 suppresses muscle differentiation at the myogenin step
in response to genotoxic stress. Cell Death Differ. 22, 560–573.
Yin, J., Sheng, B., Qiu, Y., Yang, K., Xiao, W., and Yang, H. (2016).
Role of AhR in positive regulation of cell proliferation and survival.
Cell Prolif. https://doi.org/10.1111/cpr.12282.
Yoshida, N., Yoshida, S., Koishi, K., Masuda, K., and Nabeshima, Y.
(1998). Cell heterogeneity upon myogenic differentiation: down-
regulation of MyoD and Myf-5 generates ‘‘reserve cells’’. J. Cell
Sci. 111 (Pt 6), 769–779.
Yue, F., Bi, P., Wang, C., Li, J., Liu, X., and Kuang, S. (2016). Condi-
tional loss of Pten in myogenic progenitors leads to postnatal skel-
etal muscle hypertrophy but age-dependent exhaustion of satellite
cells. Cell Rep. 17, 2340–2353.
Stem Cell Reports jVol. 10 j970–983 jMarch 13, 2018 983
Stem Cell Reports, Volume 10
Supplemental Information
The Satellite Cell Niche Regulates the Balance between Myoblast Differ-
entiation and Self-Renewal via p53
Valentina Flamini, Rachel S. Ghadiali, Philipp Antczak, Amy Rothwell, Jeremy E.
Turnbull, and Addolorata Pisconti
!
1
The satellite cell niche regulates the balance between myoblast
differentiation and self-renewal via p53
Valentina Flamini, Rachel S. Ghadiali, Philipp Antczak, Amy Rothwell, Jeremy E.
Turnbull and Addolorata Pisconti
Supplementary Information
!
2
SUPPLEMENTARY METHODS
Myofibre cultures
Mice were culled under a CO2 raising gradient and the gastrocnemius muscles
dissected, incubated with 400 U/mL collagenase type I (Worthington) for 90 min then
individual myofibres were picked using a glass Pasteur pipette whose tip had been
flame-polished and flushed with fresh primary myoblast growth medium (F12 + 0.4
mM CaCl2 + 15% horse serum + 1% penicillin + 1% streptomycin + 2 mM
GlutaMax) in a cell culture incubator (humidified 37 °C, atmospheric O2 and 5%
CO2). After 3 subsequent passages in fresh primary myoblast growth medium to
remove debris and contaminant fibroblasts, FGF2 was added at the final concentration
of 2 nM in the last passage. 2nM FGF2 was added again at 24 h and then the medium
was changed once at 48 hours supplementing again with 2nM FGF2. Although it is
still debated whether SCs derived from different myofibre types are also different and
retain these different properties once cultured ex vivo, we chose to use the
gastrocnemius muscle for myofibre preparations, rather than the most commonly used
extensor digitorum longus (EDL) muscle, because the gastrocnemius muscle contains
all myofibre types and therefore a myofibre-associated myoblast preparation obtained
from the gastrocnemius muscle is more representative of all possible “types” of SCs
than if the preparation was obtained from the EDL muscle, which contains almost
exclusively fast twitch myofibres. This was especially important because the
dispersed myoblast preparations were obtained from whole limb muscles and
therefore they too contained all “types” of SCs. The same batch of horse serum was
used throughout the work presented in this manuscript.
!
3
Myofibre-associated myoblast RNA quality control
We ran preliminary experiments to test for the presence of RNA from myofibre debris
in our myofibre-associated myoblast preparations. We re-suspended the final pellet
obtained after myoblast stripping into primary myoblast growth medium, plated this
cell suspension on gelatin-coated dishes and the day after we extracted RNA from
both the adherent cells and the supernatant. If significant amounts of contaminant
RNA from the myofibres had been present in the final myoblast pellet, we should
have been able to extract it and detect it from the supernatant of the plated myoblasts.
Instead we only detected RNA from the adherent cell population not from the
supernatant, which, once analysed by Bioanalyzer, showed only non-RNA
contaminants (Fig. S7D-E), indicating that the final pellet obtained from myoblast
isolation from myofibres is highly pure and does not contain myofibre RNA.
Preparation of dispersed myoblasts
Mice were culled under a CO2 raising gradient and all the hind-limb muscles
dissected, finely minced with scissors and then incubated with 400 U/mL collagenase
type I for 60 minutes at 37 °C. After centrifuging at low speed (30 xg) for 3 minutes
to pellet larger tissue debris, the supernatant containing mono-nucleated cells was
collected, diluted with two volumes of primary myoblast growth medium (F12 + 0.4
mM CaCl2 + 15% horse serum + 1% penicillin + 1% streptomycin + 2 mM
GlutaMax) and then centrifuged again at 500 xg to pellet the mono-nucleated cells.
Fibroblasts were removed via two subsequent pre-plating steps (2h + 1h, each one on
gelatin-coated dishes, at humidified 37 °C, atmospheric O2 and 5% CO2). After the
second pre-plating the floating cells were plated on gelatin-coated dishes in primary
myoblast growth medium, supplemented with 2 nM FGF2 and incubated for 48 or 72
hours, with one medium change at 48 hours. This protocol yielded a preparation of
!
4
great purity (over 95% MYF5+/PAX7+ cells). For the microarray experiments, at 48
and 72 h after plating, cells where washed with phosphate buffered saline (PBS) and
lysed with RLT buffer (Qiagen) prior to RNA extraction. For the reserve cell assays,
48 h after plating cells were washed twice with F12C (F12 + 0.4 mM CaCl2 ) and then
switched to primary myoblast low serum medium (F12 + 0.4 mM CaCl2 + 3% horse
serum + 1% penicillin/streptomycin + 2 mM GlutaMax) for 3 days. After 3 days, cells
were either fixed and immunostained or “washed” with F12C and incubated again in
primary myoblast growth medium (containing 15% horse serum) for 2 days, prior to
being fixed and immunostained. The same batch of horse serum was used throughout
the work presented in this manuscript.
C2C12 myoblast cultures
C2C12 myoblasts (Yaffe and Saxel, 1977) were cultured in C2C12 growth medium
(DMEM + 10% foetal bovine serum + 2 mM L-glutamine + 1% penicillin + 1%
streptomycin) up to 90% confluence then washed twice with DMEM and switched to
C2C12 low serum medium (DMEM + 3% horse serum + 2 mM L-glutamine + 1%
penicillin + 1% streptomycin) to induce differentiation and reserve cell generation
(Yoshida et al., 1998) in the presence/absence of either 20 µM Nutlin-3 (Sigma
Aldrich) or an equal volume of vehicle (DMSO, Sigma Aldrich) for an additional 3
days. After 3 days, cells were either immediately fixed and immunostained or
“washed” with DMEM and re-incubated in C2C12 growth medium (containing 10%
foetal bovine serum) for an additional 2 days prior to being fixed and immunostained.
The same batch of foetal bovine serum was used throughout the work presented in
this manuscript.
!
5
Antibodies
For immunofluorescence, primary antibodies used were: mouse anti-Pax7 (DSHB)
1:100, mouse anti-p53 (Cell Signaling, Cat No. 2524) 1:100, rabbit anti-myogenin
(SCBT, Cat. No. sc-576) 1:500, rabbit anti-SDC3 (kindly donated by Prof Brad
Olwin, University of Colorado (Cornelison et al., 2001) 1:500, rabbit anti-KI67
(Abcam, Cat. No. ab15580) 1:400, mouse anti-myogenin (DSHB, clone F5D) 1:100,
mouse anti-myosin heavy chain (DSHB, clone MF20) 1:100, rabbit anti-MYF5
(SCBT, Cat. No. sc-302) 1:500, rabbit anti-MYOD (SCBT, Cat. No. sc-760) 1:600.
Secondary antibodies conjugated to Alexa488, Alexa555 or Alexa647 (Invitrogen)
were used at 1:500. DAPI (Invitrogen) was used at 2 µM in PBS.
For western blotting, primary antibodies used were: mouse anti-p53 (Cell Signaling,
Cat No. 2524) 1:1,000; mouse anti-GAPDH (Sigma Aldrich, Cat. No. G8795)
1:3,000; mouse anti-myogenin (DSHB, clone F5D) at 1:1,000 and mouse anti-
sarcomeric myosin (DSHB, clone MF20) at 1:3,000. Secondary antibodies were from
Santa Cruz Biotechnology and used at 1:10,000. Chemiluminescence was detected on
a LAS 4,000 (GE) image doc system.
Microscopy
Microscope images were acquired on an EVOS FL imaging system (Life
Technologies) with a 10X, a 20X or a 40X objective and consistent imaging
parameters.
Microarrays
Fluorescently labelled amplified complementary RNA (cRNA) was generated using a
Low Input Quick Amp Labelling kit, One-Colour (Agilent). The method utilises an
oligo (dT) primer bearing a T7 promoter, and MMLV-RT, producing double stranded
!
6
cDNA from mRNA; the cDNA then serves as template for in vitro transcription with
T7 RNA Polymerase, which linearly amplifies target material whilst simultaneously
incorporating cyanine 3-labelled CTP. For each labelling reaction, 70 ng of total RNA
was used as input, along with appropriately diluted Spike Mix from the One-Colour
RNA Spike-In kit (Agilent). cRNA was purified using an RNeasy Mini Kit (Qiagen)
and quantified on a NanoDrop ND-1000 Spectrophotometer version 3.3.0. For array
hybridisation, 600 ng of each cyanine 3-labelled cRNA, was combined with 5 µl of
10X Gene Expression Blocking Agent and 1 µl of 25X Fragmentation Buffer (both
from the Agilent Gene Expression Hybridisation kit) in a total volume of 25 µl.
Target mixtures were then incubated at 60 °C for 30 min to fragment the RNA to
approximately 150 nucleotides. Fragmentation was terminated by cooling on ice for 1
min followed by addition of 25 µl of 2X Hi-RPM Hybridisation Buffer (Agilent).
Agilent SurePrint 8X 60K Mouse Gene Expression microarrays (Design ID 028005)
were loaded and hybridised using Agilent hardware, namely: gasket slides, SureHyb
chambers and hybridisation oven. Hybridisation was carried out at 65 °C with 10
rotations per minute for 17 h. After this time, microarrays were washed using an
Agilent Gene Expression Wash Buffer Kit according to the manufacturer’s
instructions. Arrays were scanned immediately at a resolution of 3 µm using an
Agilent DNA Microarray Scanner to generate 20 bit tiff images. Data was extracted
and QC reports generated using Agilent Feature Extraction version 11.0.1.1. Array
quality was assessed by visual inspection of each tiff and analysis of the associated
QC report, which indicates the dynamic range of the experiment, hybridisation and
background uniformity, as well as an evaluation of metrics associated with the RNA
spike-ins added to the labelling reactions.
!
7
Transfection
C2C12 myoblast cultures that had reached 90% confluence or myofibre cultures were
washed twice with PBS and switched to OptiMem before receiving a 1:3,
volume:volume mixture of siRNA (diluted in OptiMem for a final 30 or 60
nM):Lipofectamine2000. After addition of the siRNA:Lipofectamine2000 mixture,
plates where incubated for three hours in a cell culture incubator (humidified 37 °C,
atmospheric O2 and 5% CO2), then the medium was replaced with low serum medium
that contained either 20 µM Nutlin-3 or an equal volume of DMSO and incubated for
another day (for western blot analysis of C2C12 cells) or three days (for
immunofluorescence analysis of C2C12 cells) or 2 days (for immunofluorescence
analysis of myofibre cultures). siRNAs were either control (scrambled) siRNA
(Sigma Aldrich) or anti-p53 siRNA (Sigma Aldrich, SASI_Mm02_00310137).
!
8
SUPPLEMENTARY FIGURES
Supplementary Figure S1 – supplementing Fig. 1A-B: Dispersed myoblasts
cultured on gelatin in the presence of high serum and FGF2 undergo
spontaneous differentiation. Primary SC-derived myoblasts were cultured on
gelatin-coated plates in primary myoblast growth medium supplemented with 2 nM
FGF2 for 4 days. During the first 2-3 days cells mostly proliferated, around the third
day they underwent spontaneous differentiation such that myotubes were visible at
day 4 (black arrows).
!
9
Supplementary Figure S2 – supplementing Fig. 1E: Replicate reproducibility
was excellent in both dispersed and myofibre-associated myoblast cultures,
which appeared very different in their transcriptome signature during myogenic
progression. A-B) Hierarchical clustering of the Top 200 differentially expressed
genes between 48 and 72 hours in culture in myofibre-associated (A) and dispersed
(B) myoblasts. For each condition the biological replicates clustered together
suggesting high reproducibility and high quality of the data. Moreover, as visually
shown by the heatmaps, the transcriptomic signature at 48 hours is very different from
the transcriptomic signature at 72 hours for both culture conditions. C) Gene
expression regulation between 48 and 72 hours in myofibre-associated myoblasts and
in dispersed myoblasts are different.
!
10
Supplementary Figure S3 – supplementing Fig. 3: p53 protein levels do not
significantly increase over time in primary dispersed myoblast cultures. A-B)
Primary dispersed myoblasts were cultured for up to 96 hours post-isolation, lysed at
three time points: 48, 72 and 96 hours, then total proteins extracted and analysed by
western blot to detect p53 levels. Ponceau staining was used as loading control. A
representative western blot image is shown in (A) while (B) is the quantification of
three independent experiments.
!
11
Supplementary Figure S4 – supplementing Fig. 4: siRNA-mediated p53
knockdown is highly efficient. C2C12 myoblast were cultured in growth medium
until they reached 90% confluence then transfected with either control siRNA (30
nM, lane 1; 60 nM, lane 3) or anti-p53 siRNA (30 nM, lane 2; 60 nM, lane 4), 3 h
later switched to low serum medium supplemented with either DMSO or 20 µM
Nutlin-3 and cultured for an additional 24 h prior to being lysed and analysed by blot
to reveal p53 (top gel) and GAPDH (bottom gel) abundance. The intensity of the p53
bands normalised to the intensity of the GAPDH bands are reported below the gel
images as percentage of the control band for each one of the two siRNA
concentrations tested. Since the level of knockdown was comparable between 30 nM
and 60 nM siRNA, all subsequent experiments were carried out using 30 nM siRNA.
!
12
Supplementary Figure S5 – supplementing Fig. 5: Nutlin-3 treatment of primary
dispersed myoblasts inhibits proliferation without increasing cell death. A)
Primary dispersed myoblasts were expanded for two days prior to being detached and
re-plated at clonal density for an additional three days. The day after plating (time 0
treatment) culture medium was supplemented with either DMSO or Nutlin-3 at 3
different concentrations: 5 µM, 10 µM or 20 µM. Samples were fixed either the day
after (24 h treatment) or 2 days later (48 h treatment) and the numbers of cells/clone
scored and plotted as average across 3 independent experiments ± S.E.M. At 24 h,
only 20 µM Nutlin-3 significantly reduces proliferation (* = p < 0.05), at 48 h both 10
µM and 20 µM Nutlin-3 significantly reduce proliferation (** = p < 0.01). B) Primary
dispersed myoblasts were plated and expanded for one day before being treated with
either DMSO or 20 µM Nutlin-3, then fixed either the day after (24 h treatment) or 2
!
13
days later (48 h treatment), processed for TUNEL assay (green) and counter-stained
with DAPI to detect DNA (blue). At each time point, a non-treated control was also
fixed and treated with DNAse-I before TUNEL assay to induce DNA breakage – this
is labelled as DNAse-I/positive control. Although in the positive control all cells are
TUNEL+ indicating that the TUNEL assay worked, in both the DMSO and Nutlin-3-
treated samples, at both time points, only very few TUNEL+ spots are detected,
which do not co-localise with DAPI staining (hence the lack of quantification) and
might therefore be an artifact or might be late apoptotic cells whose DNA was so
extremely fragmented that no DAPI staining was detectable any more. In any case,
there appear to be no difference between the DMSO-treated and the Nutlin-3-treated
samples.
!
14
Supplementary Figure S6 – supplementing Fig. 5 and Fig. 6: A sustained
increase in p53 levels reduces C2C12 myoblast differentiation. A-C) C2C12
myoblast cultures were grown to 90% confluence then switched to low serum
medium (DMEM + 3% horse serum + 1% pen/strep + 2 mM L-glutamine) either in
the presence of 20 µM Nutlin-3 or in the presence of an equal volume of DMSO and
lysed at the following time points: Pr (= proliferating) is the day before reaching 90%
confluence; 0 is the day when 90% confluence was reached; 1, 2 and 3 are the days
subsequent to medium switch. At each time point cells were lysed and then all time
points loaded onto a 10% SDS-PAGE for western blot analysis to reveal the
abundance of myosin heavy chain (MYHC), myogenin (MYOG), p53 and GAPDH.
Representative gels from three independent experiments are shown in (A) where to
top GAPDH gel is related to the p53 gel, while the bottom GAPDH gel is related to
!
15
the MYHC and MYOG gel. In (B-C) is the quantification of three independent
experiments carried out as in (A) where the average band intensity for p53 (B) or
MYHC (C) normalised to the band intensity for its related GAPDH at each time point
is plotted. Error bars are S.E.M. across the three independent experiments. D-E)
Immunofluorescence analysis of C2C12 myoblasts switched to low serum medium
when 90% confluence was reached and treated with either 20 µM Nutlin-3 or DMSO
for 3 days prior to fixation and immunostaining to detect MYHC (green) and DNA
(DAPI, blue) reveals that differentiation is reduced by Nutlin-3 treatment. In (D) one
representative of several (> 10) independent experiments is shown. In (D) quantitation
of three independent experiments across 10 technical replicates for each one of three
biological replicates (N = 30) as the percentage of MYHC+ cells over the total
numbers of DAPI+ cells is plotted. Error bars are S.E.M. ** = p < 0.01.
!
16
Supplementary Figure S7 – cited in Experimental Procedures: Validation of the
protocol to isolate myofibre-associated myoblasts. A) Myofibre-associated
myoblasts were stripped off myofibres using trypsin as described in the Experimental
Procedures section then cytospun and immunostained to detect the SC/myoblast
markers PAX7 and MYF5: 100% of the cells cytspun were positive for at least one of
these two markers. B-C) Isolated myofibre-associated myoblasts were tested for
viability and functionality by plating them on gelatin-coated dishes in primary
!
17
myoblast growth medium. Three days later isolated myofibre-stripped myoblasts had
extensively proliferated in culture (B) and after switching them to low serum medium
they underwent differentiation into myotubes (C). D-E) Bioanalyzer analysis of the
RNA extracted from: negative control (water, lane 2); adherent cells (lane 3);
supernatant of isolated myofibre-associated myoblasts plated on gelatin-coated dishes
immediately after isolation (lane 4). Please not the massive differences in the Y-axis
scales in the chromatograms shown in (E). This analysis reveals that no RNA was
present in the supernatant indicating that no RNA was carried over from myofibre
debris into the RNA preparations that were used for microarray analysis. F)
Bioanlyzer analysis of the RNA extracted from all the biological replicates of
myofibre-associated myoblasts at 48 hours and 72 hours post-myofibre isolation.
Only for one of the 4 replicates of myofibre-associated myoblasts at 48 h, two distinct
preparations were pooled to produce one biological replicate. In all other cases each
biological replicate corresponded to one independent preparation. G) Table reporting,
for each RNA sample used in the microarray experiment: sample name, type of
myoblast prep, system for preparation, method of quantification, concentration, RIN
where applicable and 260/280 ratio where applicable. All RNA samples were eluted
in the same volume (14 µL) therefore the concentrations reported are a direct measure
of the total amount of RNA extracted.
REFERENCES
Cornelison, D.D., Filla, M.S., Stanley, H.M., Rapraeger, A.C., Olwin, B.B., 2001.
Syndecan-3 and syndecan-4 specifically mark skeletal muscle satellite cells and
are implicated in satellite cell maintenance and muscle regeneration. Dev Biol
239, 79–94. doi:10.1006/dbio.2001.0416
Yaffe, D., Saxel, O., 1977. Serial passaging and differentiation of myogenic cells
isolated from dystrophic mouse muscle. Nature 270, 725–727.
Yoshida, N., Yoshida, S., Koishi, K., Masuda, K., Nabeshima, Y., 1998. Cell
heterogeneity upon myogenic differentiation: down-regulation of MyoD and
!
18
Myf-5 generates “reserve cells.” J Cell Sci 111 ( Pt 6), 769–779.
Supplementary Table 1 – Cited in Results, section 4: Top three GO terms that are
most significantly enriched in genes with CR > 2.
Term
p-value
Genes
Negative regulation
of transport
0.00002
Lif, Il6, Nos1, Snph, Edn1, Pkia,
Adora1, Htr2a
Striated muscle
development
0.00004
Musk, Actc1, Myl2, Tnc, Myog, Ttn,
Neurl2
Taxis
0.00005
Ccr8, Ccl2, Cxcl3, Cxcl2, Ecscr, Amot,
Ccl5, Ccl7, Slit2
Supplementary Table 2 – Cited in Results, section 4: Top three GO terms that are
most significantly enriched in genes with CR < 0.5.
Term
p-value
Genes
Fold
enrichment
Cell adhesion
0.0004
Col18a1, Selp, Pcdhb7, Pcdhb6, Pcdhb3,
Pcdhb4, Itgb4, Cdhr5, Vtn, Pcdh17,
Megf10, Dchs1, Cldn15, Aplp1, Wnt7b,
Lamb3, Itga6, Lama5, Tro, Otog, Msln
2.42
Cell morphogenesis
involved in
differentiation
0.0016
Col18a1, Ablim1, Sema6a, Notch1,
Slc1a3, Rtn4rl1, Cxcr4, Sema3a, Gas1,
Ngfr, Cxcl12
3.35
Cell fate commitment
0.0019
Notch3, Fgfr4, Notch1, Tbx2, Pax7,
Pparg, Cyp26b1, Gas1, Sox8
3.96
Supplementary Table 3 – Cited in Experimental Procedures and in Results, section
4: Comparative analysis of genes that change between 48 and 72 hours according to
the formula: [(72h/48h)dispersed]/[(72h/48h)myofibre-associated] filtered through a manually
curated list of genes involved in myogenesis.
!
19
Gene$Name$
Fold$Change$
q-value$
Diras2'
0.054013395(
0.0000000(
Mc4r'
0.093310186(
0.0000000(
Aqp5'
0.096597149(
0.0000000(
Ngfr'
0.105205253(
0.0000000(
Cd5l'
0.179134772(
0.0000000(
Sema6b'
0.228863483(
0.0000000(
Trpc3'
0.233711819(
0.0235839(
Fzd6'
0.235680061(
0.0000000(
Fgfr4'
0.236169718(
0.0000000(
Epha1'
0.250161594(
0.0000000(
Itgb4'
0.25121833(
0.0000000(
Pitx3'
0.255210552(
0.0000000(
Itga6'
0.263171022(
0.0000000(
Notch1'
0.277726352(
0.0000000(
Col18a1'
0.278990429(
0.0000000(
Tmem119'
0.290233444(
0.0000000(
Notch3'
0.291949448(
0.0000000(
Col10a1'
0.300396372(
0.0000000(
Cxcr4'
0.315090189(
0.0000000(
Cdkn2a'
0.316454481(
0.0000000(
Ly6a'
0.316960439(
0.0000000(
Sox8'
0.323270186(
0.0000000(
Tmem44'
0.328784523(
0.0000000(
Megf10'
0.335582833(
0.0000000(
Cdc42ep1'
0.337906165(
0.0000000(
Hes6'
0.338134183(
0.0000000(
Pax7'
0.344352698(
0.0000000(
Hey1'
0.345761151(
0.0000000(
Cxcl12'
0.348607325(
0.0000000(
Sema6a'
0.349215321(
0.0000000(
Rnf152'
0.350092944(
0.0000000(
Tmem191c'
0.350341532(
0.0235839(
Pparg'
0.361863203(
0.0000000(
Rnf144a'
0.400662004(
0.0000000(
Cdkn2b'
0.401434376(
0.0000000(
Lama5'
0.412932128(
0.0000000(
Sema3a'
0.41859219(
0.0000000(
Tcf7'
0.423415284(
0.0000000(
Pdgfc'
0.427927325(
0.0235839(
Tmem176a'
0.428260942(
0.0235839(
Jam2'
0.43211217(
0.0000000(
Pdgfrl'
0.436975917(
0.0000000(
Sdc3'
0.437947492(
0.0000000(
!
20
Timp4'
0.442403723(
0.0000000(
Cnnm2'
0.442786114(
0.0000000(
Mmp15'
0.449408686(
0.0000000(
Rnf125'
0.454254155(
0.0000000(
Lamb3'
0.455321565(
0.0000000(
Rnf208'
0.463874638(
0.0000000(
Tnfaip8l1'
0.465211468(
0.0000000(
Tmem30b'
0.471209484(
0.0235839(
Cd200'
0.472625308(
0.0000000(
Fzd2'
0.481798531(
0.0000000(
Ace'
0.485118212(
0.0000000(
Lrp12'
0.487980149(
0.0000000(
Wnt7b'
0.491717209(
0.0000000(
Gper'
0.493160426(
0.0000000(
Creb3l1'
0.50032382(
0.0000000(
Tcf4'
0.502701894(
0.0000000(
Tnfrsf25'
0.511265133(
0.0235839(
Trim62'
0.511388518(
0.0000000(
Megf9'
0.512288348(
0.0235839(
Tmem176b'
0.512579672(
0.0335022(
Chrna1'
0.512895398(
0.0000000(
Ephb4'
0.513271937(
0.0000000(
Cdh15'
0.513597528(
0.0000000(
Igfbp4'
0.518688547(
0.0000000(
Megf6'
0.520352228(
0.0335022(
Tmem195'
0.524521151(
0.0000000(
Cd276'
0.526542761(
0.0000000(
Traf5'
0.534764828(
0.0235839(
Map3k11'
0.541734759(
0.0000000(
Cd2ap'
0.550591348(
0.0000000(
Tmem121'
0.56017007(
0.0235839(
Smad6'
0.56064096(
0.0000000(
Tmem9'
0.562422561(
0.0000000(
Il20rb'
0.563193529(
0.0000000(
Tmem229b'
0.571979361(
0.0000000(
Wnt6'
0.579395389(
0.0000000(
Ctnnb1'
0.58218307(
0.0235839(
Pth1r'
0.583389597(
0.0000000(
Hspa12b'
0.58409558(
0.0000000(
Il16'
0.587001345(
0.0235839(
Cdc42ep4'
0.591780882(
0.0000000(
Cd97'
0.595509301(
0.0000000(
Jak3'
0.600945954(
0.0000000(
Cd38'
0.603198555(
0.0235839(
Tmem86a'
0.606710437(
0.0000000(
!
21
Il18rap'
0.608983054(
0.0335022(
Mmp11'
0.61109568(
0.0000000(
F2r'
0.618226226(
0.0000000(
Cdk19'
0.618693237(
0.0000000(
Il34'
0.619225504(
0.0000000(
Mmp17'
0.62185378(
0.0235839(
Map4k2'
0.623623769(
0.0000000(
Cxcr7'
0.624860007(
0.0000000(
Hmgb3'
0.624875968(
0.0000000(
Tmem173'
0.630723339(
0.0000000(
Six2'
0.631694902(
0.0335022(
Ctnnbip1'
0.63742738(
0.0335022(
Fzd3'
0.639107577(
0.0335022(
Socs2'
0.64750778(
0.0335022(
Hmgn5'
0.648216068(
0.0235839(
Stat2'
0.658367534(
0.0235839(
Cdk2ap2'
0.659850247(
0.0000000(
Met'
0.665855282(
0.0000000(
Sdc4'
0.667108623(
0.0235839(
Tcf3'
0.668449518(
0.0000000(
Trim12'
0.671382361(
0.0335022(
Ilkap'
0.675550964(
0.0335022(
Ccna2'
0.675932308(
0.0235839(
Mapk3'
0.686093157(
0.0000000(
Itga3'
0.688256775(
0.0000000(
Ilf2'
0.689515315(
0.0000000(
Grb10'
0.691704377(
0.0000000(
Tmem110'
0.694182038(
0.0335022(
Itm2c'
0.701841327(
0.0335022(
Cd3eap'
0.704417076(
0.0335022(
Rnf38'
0.704939689(
0.0335022(
Nfkbil2'
0.705436177(
0.0335022(
Map3k12'
0.711453353(
0.0235839(
Mybl2'
0.712538267(
0.0235839(
Cdh8'
0.733437491(
0.0235839(
Cdkn2c'
0.733704766(
0.0235839(
Cd151'
0.734003656(
0.0235839(
Casp2'
0.740118071(
0.0335022(
Cdk10'
0.742337344(
0.0235839(
Atoh8'
0.743759899(
0.0235839(
Cdk2ap1'
0.759884675(
0.0335022(
Tmem179b'
0.790983217(
0.0335022(
Tmem50a'
0.791326965(
0.0335022(
Col9a2'
0.805013649(
0.0235839(
Nos2'
0.808038002(
0.0235839(
!
22
Rnf160'
1.182986633(
0.0335022(
Ppara'
1.226111384(
0.0235839(
Rnf103'
1.265119029(
0.0000000(
Abl2'
1.28007608(
0.0235839(
Rnf181'
1.28600617(
0.0335022(
Sod2'
1.288438354(
0.0235839(
Tmem126a'
1.297795278(
0.0335022(
Large'
1.304309246(
0.0335022(
Tmem9b'
1.31584281(
0.0335022(
Rnf170'
1.32832226(
0.0235839(
Lamc1'
1.338170117(
0.0335022(
Crybg3'
1.352691383(
0.0235839(
Rnf19b'
1.377284374(
0.0335022(
Nfix'
1.428716223(
0.0335022(
Cd300lg'
1.487985697(
0.0335022(
Itga2b'
1.518339592(
0.0335022(
Cdc42ep3'
1.520005509(
0.0335022(
Trim23'
1.53151972(
0.0335022(
Cdkn1c'
1.568324588(
0.0235839(
Ccng1'
1.608196047(
0.0335022(
Tmem143'
1.681525317(
0.0000000(
Ar'
1.706092125(
0.0235839(
Mef2d'
1.71779398(
0.0000000(
Tmem51'
1.720317928(
0.0335022(
Myc'
1.727137888(
0.0235839(
Tmem14a'
1.741341997(
0.0000000(
Rnf113a2'
1.748848333(
0.0000000(
Traf3'
1.749643353(
0.0000000(
Rnf128'
1.758049346(
0.0000000(
Col4a5'
1.764058974(
0.0235839(
Hes2'
1.807606777(
0.0335022(
Rnf123'
1.822805482(
0.0335022(
Itga5'
1.824026319(
0.0335022(
Tmem65'
1.844739221(
0.0000000(
Mapk14'
1.877026513(
0.0000000(
Tmem25'
1.877475243(
0.0000000(
Tmem170'
1.890353302(
0.0335022(
Myh7'
1.904240751(
0.0335022(
Hdac9'
2.039262001(
0.0335022(
Cd93'
2.077909983(
0.0335022(
Ang2'
2.119929417(
0.0000000(
Fgf10'
2.136513435(
0.0335022(
Rnf39'
2.142887692(
0.0235839(
Rnf115'
2.185271495(
0.0335022(
Tmem74'
2.189673882(
0.0235839(
!
23
Wnt5a'
2.225643565(
0.0000000(
Tgfbr3'
2.243237532(
0.0335022(
Nos1'
2.244337119(
0.0000000(
Tnfaip3'
2.370856078(
0.0335022(
Tmem22'
2.382163203(
0.0335022(
Timp3'
2.389738255(
0.0000000(
Grb14'
2.450073063(
0.0335022(
Icam1'
2.480875001(
0.0000000(
Cd34'
2.502741332(
0.0235839(
Fgf13'
2.522457048(
0.0335022(
Tnfrsf26'
2.526705886(
0.0335022(
Cxcl1'
2.571387986(
0.0335022(
Rnf150'
2.613735241(
0.0000000(
Cxcr6'
2.666414714(
0.0235839(
Hspb1'
2.671942324(
0.0335022(
Igf2bp1'
2.685073278(
0.0335022(
Lama4'
2.694176556(
0.0000000(
Col11a1'
2.81007029(
0.0000000(
Il1rl1'
2.883004106(
0.0235839(
Angpt2'
2.889406621(
0.0000000(
Ctgf'
2.917124574(
0.0000000(
Trim55'
2.943127311(
0.0235839(
Sema3c'
2.962051241(
0.0000000(
Itga11'
3.097928862(
0.0335022(
Smad3'
3.143103306(
0.0335022(
Itgb7'
3.180213711(
0.0000000(
Des'
3.276458094(
0.0335022(
Wnt9a'
3.473096788(
0.0000000(
Tmem132e'
3.637282601(
0.0000000(
Tnfaip6'
3.705210584(
0.0000000(
Crym'
3.715182105(
0.0335022(
Pdgfra'
3.799874363(
0.0000000(
Tmem56'
3.824028259(
0.0000000(
Stac3'
4.096755699(
0.0000000(
Itga2'
4.151801886(
0.0000000(
Myl2'
4.478819312(
0.0000000(
F3'
4.53209333(
0.0000000(
Itgb1bp2'
5.052382919(
0.0000000(
Cxcl2'
5.21740209(
0.0000000(
Il6'
5.883417321(
0.0000000(
Cxcl5'
7.072889179(
0.0335022(
Mmp10'
8.091003904(
0.0000000(
Mmp3'
9.798800323(
0.0000000(
Csf2'
10.41366572(
0.0000000(
Postn'
13.13746879(
0.0000000(
!
24
Fgf21'
13.98192167(
0.0000000(
Cxcl3'
15.22060208(
0.0000000(
Myl3'
18.83977572(
0.0000000(
Myog'
20.03261426(
0.0000000(
Mmp13'
28.20644517(
0.0000000(
!
!
!

File (1)

ResearchGate has not been able to resolve any citations for this publication.
Article
Full-text available
Skeletal muscle stem cells (satellite cells [SCs]) are normally maintained in a quiescent (G0) state. Muscle injury not only activates SCs locally, but also alerts SCs in distant uninjured muscles via circulating factors. The resulting GAlert SCs are adapted to regenerative cues and regenerate injured muscles more efficiently, but whether they provide any long-term benefits to SCs is unknown. Here, we report that embryonic myogenic progenitors lacking the phosphatase and tensin homolog (Pten) exhibit enhanced proliferation and differentiation, resulting in muscle hypertrophy but fewer SCs in adult muscles. Interestingly, Pten null SCs are predominantly in the GAlert state, even in the absence of an injury. The GAlert SCs are deficient in self-renewal and subjected to accelerated depletion during regeneration and aging and fail to repair muscle injury in old mice. Our findings demonstrate a key requirement of Pten in G0 entry of SCs and provide functional evidence that prolonged GAlert leads to stem cell depletion and regenerative failure.
Article
Full-text available
Background The skeletal muscle stem cell niche provides an environment that maintains quiescent satellite cells, required for skeletal muscle homeostasis and regeneration. Syndecan-3, a transmembrane proteoglycan expressed in satellite cells, supports communication with the niche, providing cell interactions and signals to maintain quiescent satellite cells. Results Syndecan-3 ablation unexpectedly improves regeneration in repeatedly injured muscle and in dystrophic mice, accompanied by the persistence of sublaminar and interstitial, proliferating myoblasts. Additionally, muscle aging is improved in syndecan-3 null mice. Since syndecan-3 null myofiber-associated satellite cells downregulate Pax7 and migrate away from the niche more readily than wild type cells, syxndecan-3 appears to regulate satellite cell homeostasis and satellite cell homing to the niche. Conclusions Manipulating syndecan-3 provides a promising target for development of therapies to enhance muscle regeneration in muscular dystrophies and in aged muscle. Electronic supplementary material The online version of this article (doi:10.1186/s13395-016-0104-8) contains supplementary material, which is available to authorized users.
Article
Full-text available
Satellite cells (SCs) are skeletal muscle stem cells residing quiescent around healthy muscle fibres. In response to injury or disease SCs activate, proliferate and eventually differentiate and fuse to one another to form new muscle fibres, or to existing damaged fibres to repair them. The sulfated polysaccharide heparan sulfate (HS) is a highly variable biomolecule known to play key roles in the regulation of cell fate decisions, though the changes that muscle HS undergoes during SC differentiation are unknown. Here we show that the sulfation levels of HS increase during SC differentiation; more specifically, we observe an increase in 6-O and 2-O-sulfation in N-acetylated disaccharides. Interestingly, a specific increase in 6-O sulfation is also observed in the heparanome of ageing muscle, which we show leads to promotion of FGF2 signalling and satellite cell proliferation, suggesting a role for the heparanome dynamics in age-associated loss of quiescence. Addition of HS mimetics to differentiating SC cultures results in differential effects: an oversulfated HS mimetic increases differentiation and inhibits FGF2 signalling, a known major promoter of SC proliferation and inhibitor of differentiation. In contrast, FGF2 signalling is promoted by an N-acetylated HS mimetic, which inhibits differentiation and promotes SC expansion. We conclude that the heparanome of SCs is dynamically regulated during muscle differentiation and ageing, and that such changes might account for some of the phenotypes and signalling events that are associated with these processes.
Article
Full-text available
In Duchenne muscular dystrophy, progressive loss of muscle tissue is accompanied by fibrosis, chronic inflammation and reduced muscle regenerative capacity. Although much is known about the development of fibrosis and chronic inflammation in muscular dystrophy, less is known about how they are mechanistically linked to loss of muscle regenerative capacity. We have developed a proteomics method to discover dystrophy-associated changes in the muscle progenitor cell niche, which identified serine proteases, and especially neutrophil elastase, as candidates. We show that elastase activity is increased in dystrophic (mdx4cv) muscle and impairs myoblast survival in culture. While the effect of elastase on C2C12 cell survival correlates with the kinetics of elastase-mediated degradation of the substrate to which the cells adhere, the effect of elastase on satellite cell-derived primary myoblast growth and differentiation is substrate-independent and even more dramatic than the effect on C2C12 cells, suggesting a detrimental role for elastase on myogenesis in vivo. Additionally, elastase impairs differentiation of both primary and C2C12 myoblasts into myotubes. Our findings evidence the importance of neutrophil-mediated inflammation in muscular dystrophy and indicate elastase-mediated regulation of myoblast behaviour as a potential mechanism underlying loss of regenerative capacity in dystrophic muscle.
Article
Full-text available
Satellite cells are muscle-resident stem cells capable of self-renewal and differentiation to repair injured muscles. However, muscle injury often leads to an ischemic hypoxia environment that impedes satellite cell differentiation and reduces the efficiency of muscle regeneration. Here we performed microarray analyses and identified the basic helix-loop-helix family transcription factor Bhlhe40 as a candidate mediator of the myogenic inhibitory effect of hypoxia. Bhlhe40 is strongly induced by hypoxia in satellite cell-derived primary myoblasts. Overexpression of Bhlhe40 inhibits Myog expression and mimics the effect of hypoxia on myogenesis. Inhibition of Bhlhe40, conversely, up-regulates Myog expression and promotes myogenic differentiation. Importantly, Bhlhe40 knockdown rescues myogenic differentiation under hypoxia. Mechanistically, Bhlhe40 binds to the proximal E-boxes of the Myog promoter and reduces the binding affinity and transcriptional activity of MyoD on Myog. Interestingly, hypoxia induces Bhlhe40 expression independent of HIF1α but through a novel p53-dependent signaling pathway. Our study establishes a crucial role of Bhlhe40 in mediating the repressive effect of hypoxia on myogenic differentiation and suggests that inhibition of Bhlhe40 or p53 may facilitate muscle regeneration after ischemic injuries.
Article
Full-text available
Muscle stem cells, termed satellite cells, are crucial for skeletal muscle growth and regeneration. In healthy adult muscle, satellite cells are quiescent but poised for activation. During muscle regeneration, activated satellite cells transiently re-enter the cell cycle to proliferate and subsequently exit the cell cycle to differentiate or self-renew. Recent studies have demonstrated that satellite cells are heterogeneous and that subpopulations of satellite stem cells are able to perform asymmetric divisions to generate myogenic progenitors or symmetric divisions to expand the satellite cell pool. Thus, a complex balance between extrinsic cues and intrinsic regulatory mechanisms is needed to tightly control satellite cell cycle progression and cell fate determination. Defects in satellite cell regulation or in their niche, as observed in degenerative conditions such as aging, can impair muscle regeneration. Here, we review recent discoveries of the intrinsic and extrinsic factors that regulate satellite cell behaviour in regenerating and degenerating muscles. © 2015. Published by The Company of Biologists Ltd.
Article
Full-text available
Acute muscle injury and physiological stress from chronic muscle diseases and aging lead to impairment of skeletal muscle function. This raises the question of whether p53, a cellular stress sensor, regulates muscle tissue repair under stress conditions. By investigating muscle differentiation in the presence of genotoxic stress, we discovered that p53 binds directly to the myogenin promoter and represses transcription of myogenin, a member of the MyoD family of transcription factors that plays a critical role in driving terminal muscle differentiation. This reduction of myogenin protein is observed in G1-arrested cells and leads to decreased expression of late but not early differentiation markers. In response to acute genotoxic stress, p53-mediated repression of myogenin reduces post-mitotic nuclear abnormalities in terminally differentiated cells. This study reveals a mechanistic link previously unknown between p53 and muscle differentiation, and suggests new avenues for managing p53-mediated stress responses in chronic muscle diseases or during muscle aging.Cell Death and Differentiation advance online publication, 12 December 2014; doi:10.1038/cdd.2014.189.
Chapter
The regulation of stem cells that maintain and regenerate postnatal tissues depends on extrinsic signals originating from their microenvironment, commonly referred to as the stem cell niche. Complex higher-order regulatory interrelationships with the tissue and factors in the systemic circulation are integrated and propagated to the stem cells through the niche. The stem cell niche in skeletal muscle tissue is both a paradigm for a structurally and functionally relatively static niche that maintains stem cell quiescence during tissue homeostasis, and a highly dynamic regenerative niche that is subject to extensive structural remodeling and a flux of different support cell populations. Conditions ranging from aging to chronically degenerative skeletal muscle diseases affect the composition of the niche and thereby impair the regenerative potential of muscle stem cells. A holistic and integrative understanding of the extrinsic mechanisms regulating muscle stem cells in health and disease in a broad systemic context will be imperative for the identification of regulatory hubs in the niche interactome that can be targeted to maintain, restore, or enhance the regenerative capacity of muscle tissue. Here, we review the microenvironmental regulation of muscle stem cells, summarize how niche dysfunction can contribute to disease, and discuss emerging therapeutic implications.
Article
The aryl hydrocarbon receptor (AhR) is an important nuclear transcription factor that is best known for mediating toxic responses by adjusting numbers of metabolism-related enzymes, including CYP1A1 and CYP1B1. Previous findings have revealed that, in addition to negatively regulating cell proliferation and survival, AhR may also positively regulate these pathways. Here, we review these findings and summarize distinct mechanisms by which AhR promotes cell proliferation and survival, including modulation of receptor expression, growth factor signalling and apoptosis, regulating the cell cycle and promoting cytokine expression. This review will aid better understanding the role of AhR in positive regulation of cell proliferation and survival.