ArticlePDF Available

Abstract and Figures

microRNAs (miRs) control stem cell biology and fate. Ubiquitously expressed and conserved miR-16 was the first miR implicated in tumorigenesis. miR-16 is low in muscle during developmental hypertrophy and regeneration. It is enriched in proliferating myogenic progenitor cells but is repressed during differentiation. The induction of miR-16 blocks myoblast differentiation and myotube formation while knockdown enhances it. Despite a central role for miR-16 in myogenic cell biology, how it mediates its potent effects is incompletely defined. In this investigation, global transcriptomic and proteomic analyses after miR-16 knockdown in proliferating C2C12 myoblasts revealed how miR-16 influences myogenic cell fate. Eighteen hours after miR-16 inhibition, ribosomal protein gene expression levels were higher relative to control myoblasts and p53 pathway-related gene abundance was lower. At the protein level at this same timepoint, miR-16 knockdown globally upregulated TCA cycle proteins while downregulating RNA metabolism-related proteins. miR-16 inhibition induced specific proteins associated with myogenic differentiation such as ACTA2, EEF1A2, and OPA1. We extend prior work in hypertrophic muscle tissue and show that miR-16 is lower in mechanically overloaded muscle in vivo. Our data collectively point to how miR-16 is implicated in aspects of myogenic cell differentiation. A deeper understanding of the role of miR-16 in myogenic cells has consequences for muscle developmental growth, exercise-induced hypertrophy, and regenerative repair after injury, all of which involve myogenic progenitors.
Content may be subject to copyright.
RAPID REPORT
Musculoskeletal Biology and Bioengineering
MicroRNA control of the myogenic cell transcriptome and proteome: the role of
miR-16
Seongkyun Lim,
1
David E. Lee,
1
Francielly Morena da Silva,
1
Pieter J. Koopmans,
1,2
Ivan J. Vechetti Jr,
3
Ferdinand von Walden,
4
Nicholas P. Greene,
1,2
and Kevin A. Murach
1,2
1
Department of Health, Human Performance, and Recreation, Exercise Science Research Center, University of Arkansas,
Fayetteville, Arkansas, United States;
2
Cell and Molecular Biology Graduate Program, University of Arkansas, Fayetteville,
Arkansas, United States;
3
Department of Nutrition and Health Sciences, University of Nebraska-Lincoln, Lincoln, Nebraska, United
States; and
4
Neuropediatrics, Department of Womens and Childrens Health, Karolinska Institutet, Stockholm, Sweden
Abstract
MicroRNAs (miRs) control stem cell biology and fate. Ubiquitously expressed and conserved miR-16 was the rst miR implicated
in tumorigenesis. miR-16 is low in muscle during developmental hypertrophy and regeneration. It is enriched in proliferating myo-
genic progenitor cells but is repressed during differentiation. The induction of miR-16 blocks myoblast differentiation and myo-
tube formation, whereas knockdown enhances these processes. Despite a central role for miR-16 in myogenic cell biology, how
it mediates its potent effects is incompletely dened. In this investigation, global transcriptomic and proteomic analyses after
miR-16 knockdown in proliferating C2C12 myoblasts revealed how miR-16 inuences myogenic cell fate. Eighteen hours after
miR-16 inhibition, ribosomal protein gene expression levels were higher relative to control myoblasts and p53 pathway-related
gene abundance was lower. At the protein level at this same time point, miR-16 knockdown globally upregulated tricarboxylic acid
(TCA) cycle proteins while downregulating RNA metabolism-related proteins. miR-16 inhibition induced specic proteins associated
with myogenic differentiation such as ACTA2, EEF1A2, and OPA1. We extend prior work in hypertrophic muscle tissue and show that
miR-16 is lower in mechanically overloaded muscle in vivo. Our data collectively point to how miR-16 is implicated in aspects of myo-
genic cell differentiation. A deeper understanding of the role of miR-16 in myogenic cells has consequences for muscle developmen-
tal growth, exercise-induced hypertrophy, and regenerative repair after injury, all of which involve myogenic progenitors.
lncRNA; proteomics; RNA-sequencing; satellite cells; skeletal muscle
INTRODUCTION
MicroRNAs (miRNAs) destabilize mRNAs and/or prevent
their translation and are key regulators of myogenic stem cell
(satellite cell) biology (1). These small noncoding RNAs control
nearly every aspect of myogenic cell function including main-
tenance of quiescence (25), activation (6,7), proliferation
(8,9), self-renewal (10), migration (11), and differentiation (8,
12,13). The depletion of miRNAs from myogenic cells results
in swift cell death (2,8,14). Dysregulation of miRNAs in satel-
lite cells in vivo may negatively affect muscle development,
adaptability to exercise, the muscle microenvironment with
aging, and regenerative capacity after injury; all these proc-
esses depend on satellite cells (1524). Furthermore, miRNAs
are released from satellite cells in extracellular vesicles (EVs)
and taken up by recipient cells throughout muscle in vivo (14,
25,26). Satellite cell communication throughout muscle via
miRNAs in EVs contributes to muscle adaptation during
loading (27,28). Understanding the role of miRNAs in myo-
genic progenitors may accelerate the discovery of miRNA-
based therapeutics that affect muscle plasticity via fusion-de-
pendent and/or -independent mechanisms (27,29).
miR-16 is highly conserved and ubiquitously expressed (30
32). It was the rst miRNA implicated in tumorigenesis and is a
highly inuential miRNA in numerous cellular processes, spe-
cically as it relates to tumor growth (32,33). In skeletal muscle
tissue that contains both mature muscle bers and myogenic
satellite cells, miR-16 content rises during avian embryonic de-
velopment but begins to wane at birth concomitant with sec-
ondary myogenesis (34). Low miR-16 levels during development
areassociatedwithmusclehypertrophy(35,36). We reported
that muscle miR-16 levels decline after resistance exercise fol-
lowing a week of acclimation to training and that knockdown in
myotubes increases muscle protein synthesis (37). Skeletal mus-
cle miR-16 is also lower in regenerating relative to uninjured
muscle (6). The repression of miR-16 in muscle is therefore
Correspondence: N. P. Greene (npgreene@uark.edu); K. A. Murach (kmurach@uark.edu).
Submitted 24 February 2023 / Revised 20 March 2023 / Accepted 20 March 2023
http://www.ajpcell.org 0363-6143/23 Copyright ©2023 the American Physiological Society. C1101
Am J Physiol Cell Physiol 324: C1101C1109, 2023.
First published March 27, 2023; doi:10.1152/ajpcell.00071.2023
Downloaded from journals.physiology.org/journal/ajpcell at Univ of Arkansas Fayetteville (130.184.252.076) on May 18, 2023.
associated with postnatal myober growth and regeneration as
well as the exercise-induced hypertrophic response. In myo-
genic precursor cells, miR-16 is induced upon activation (2,6,
38) and enriched during proliferation (2,14)butgradually
declines in culture (2) and with differentiation (34,36,39). Some
evidence suggests that miR-16-5p is high in quiescent satellite
cells but still drops precipitously during differentiation (1). miR-
16 overexpression prevents myogenic cell differentiation and
myotube formation in vitro (34), whereas knockdown enhances
these processes (34,40). MyoD contributes to myogenic cell dif-
ferentiation (4148). miR-16 knockdown in murine myober-
associated myogenic cell culture may increase the proportion of
MyoD þsatellite cells by 3 days but reduces it by 5 days, point-
ing to an effect on myogenic cell behavior and fate (6). The liter-
ature collectively points to miR-16 having a key function in
myogenic cells and in determining skeletal muscle mass.
Despite its importance, how miR-16 exerts its effects in myo-
genic cells is incompletely dened.
The purpose of this investigation was to provide detailed
information on how miR-16 inuences myogenic cell biol-
ogy. First, we evaluated miR-16 levels in mechanically over-
loaded muscle tissue in vivo, which is a well-established
model of rapid growth associated with satellite cell differen-
tiation and fusion (49). We then inhibited miR-16-5p in pro-
liferating C2C12 myoblasts and performed RNA-sequencing
(RNA-seq) and discovery proteomics. C2C12s express the sat-
ellite cell marker Pax7 (50) and can model the behaviors of
primary myogenic progenitors in vitro (5153). Our experi-
ments point to a multifaceted role for miR-16 repression in
myogenic cell differentiation. We thus shed light on poten-
tial mechanisms whereby miR-16 inuences satellite cell-
mediated muscle growth and regenerative potential in vivo
(6,36). We also provide general information on miR-16 tar-
gets, which may be informative in nonmuscle cell types
where miR-16 is enriched (30).
MATERIALS AND METHODS
Murine Experiment
Female C57BL6/J mice (23 mo of age) were utilized for in
vivo overload experiments. These experiments were per-
formed for a prior publication from our laboratory (54), and
the mice were injected with 5-ethenyl uridine 5 h before
being euthanized. Experiments were approved by the
Institutional Animal Care and Use Committee (IACUC) of
the University of Kentucky. Mice were housed in a tempera-
ture- and humidity-controlled room maintained on a 14:10-h
light-dark cycle, and food and water were provided ad libi-
tum throughout experimentation. Briey, synergist ablation
surgery to overload the plantaris muscle was performed as
described by Murach et al. (54). Under isourane anesthesia,
a portion of the gastrocnemius and soleus was removed,
leaving the plantaris to be mechanically overloaded during
reambulation (n= 4). Sham-operated mice (no removal of
muscle) were controls (n=3).Seventy-twohoursaftersur-
gery, animals were euthanized in the morning via a lethal
dosage of pentobarbital sodium injected intraperitoneally
followed by cervical dislocation. Plantaris muscle tissue
was harvested and frozen in liquid nitrogen and stored for
downstream analyses. miR-16 level in tissue was evaluated
with an unpaired directional ttest in GraphPad Prism
(Boston, MA).
Cell Culture Experiments
C2C12 murine myoblasts (CRL-1772, ATCC, Manassas, VA)
were plated in six-well plates (5 10
4
cells/well) with 2 mL of
Dulbeccosmodied Eagle medium (DMEM) combined with
20% fetal bovine serum (FBS) and 1% penicillin-streptomycin
(P/S). Cells were incubated at 37Cwith5%CO
2
,andmedium
was changed every 48 h. Upon 4050% conuence, cells were
washed with PBS for transfection (37). For proteomics ve
technical replicates were utilized, and for transcriptomics
three control and two knockdown wells were employed. miR-
16 levels in cells were evaluated in Prism with an unpaired
directional ttest with n= 5 technical replicates.
Transfection of C2C12 Myoblasts
Plasmids encoding either an empty vector control
(pCMVmiR;PCMVMIR,Origene,Rockville,MD)ormiR-16-1
(miR-16-5p) inhibitor were transferred into DH5-a Escherichia
coli as we have previously described (37). Plasmid DNA was
ampliedandisolatedfrombacteriawiththePureLinkHiPure
Plasmid Filter Maxiprep kit (K211017, Life Technologies,
Carlsbad,CA).FormiR-16inhibition,weutilizedanti-miRin-
hibitor (AM17000, Ambion, Austin, TX). Plasmid DNA (1 μg)
was diluted in 50 μL of Opti-MEM reduced serum medium
(31985088, Life Technologies), combined with 4 μLof
Lipofectamine 2000 (11668019, Life Technologies) diluted in
50 μL of Opti-MEM, and incubated for 20 min to allow lipid/
DNA complexes to form. Medium was replaced with Opti-
MEM, and lipid/DNA complexes were added and incubated
for 5 h at 37Cwith5%CO
2
. After 5-h incubation, medium
was replaced with normal growth medium for 18 h. The ef-
ciency of the inhibition of miR-16 was tested and validated
previously (37).
RNA Isolation, Quality Check, and qPCR for miR-16
Myoblast and muscle tissue RNA were isolated with
TRIzol reagent, which was then homogenized with a
Polytron or bullet homogenizer. After homogenization, RNA
was isolated with the phase separation method by addition
of chloroform or bromochloropropane followed by centrifu-
gation. The aqueous phase was transferred to a new 1.5-mL
tube, and an equal amount of 70% of diethyl pyrocarbonate
(DEPC)-treated ethanol was added. RNA isolation was fur-
ther processed with the Invitrogen RNA isolation kit
(K145002, Invitrogen, Carlsbad, CA) or the Zymo Direct-zol
kit (Zymo Research, Irvine, CA). RNA concentration was
determined with a BioTek Take3 microplate with a BioTek
PowerWave XS microplate reader as previously described
(55) or NanoDrop (ThermoFisher Scientic). RNA samples
were only accepted if 260 nm-to-280 nm ratio was >2.0.
Samples were stored at 80Cuntilfurtheruse.
Reverse transcription (RT) of miR-16 was performed with
the TaqMan MicroRNA Reverse Transcription Kit (4366596,
Applied Biosystems, Waltham, MA). Briey, 200 ng of total
miRNA was added to a master mix comprised of 1 lLofa
10TaqMan probe (PN4427975, Applied Biosystems) for U6
and miR-16 each (RT:001973 and RT:000391, respectively,
Applied Biosystems), 0.3 lLofdNTPs,3lLofMultiScribe
miR-16 IN MYOGENIC CELLS
C1102 AJP-Cell Physiol doi:10.1152/ajpcell.00071.2023 www.ajpcell.org
Downloaded from journals.physiology.org/journal/ajpcell at Univ of Arkansas Fayetteville (130.184.252.076) on May 18, 2023.
RTase, 1.5 lLof10RTase buffer, 0.19 lLofRNaseinhibitor,
and nuclease-free water added up to 15 lLoftotalvolume.RT
reaction occurred with a hold step of 30 min at 16C, followed
by 30 min at 42Candnally 5 min at 85C. Samples were held
at 4C until further analysis. RT-qPCR was performed with the
QuantStudio 3 Real-Time PCR system (Applied Biosystems). A
20-lL reaction composed of an adequate amount of TaqMan
probes plus TaqMan Fast Advanced Master Mix (4444556,
Applied Biosystems) was used to amplify cDNA. Samples
followed a protocol consisting of incubation at 95Cfor4
min, followed by 45 cycles of denaturation, annealing, and
extension at 95Cand60
C. TaqMan probes were meas-
ured at the end of the extension step of each cycle.
Fluorescence-labeled 20TaqMan probes (PN4427975,
Applied Biosystems) included U6 and miR-16 (TM:001973 and
TM:000391, respectively, Applied Biosystems). Results were
analyzed with QuantStudio Software. Cycle threshold (C
T
)
was determined, and the DC
T
value was calculated as the dif-
ference between C
T
value and U6 C
T
value. U6 C
T
values were
not different between experimental conditions. Final quanti-
cation of gene expression was calculated with the DDC
T
method. Relative quantication was calculated as 2DDCTand
expressed as arbitrary units.
RNA-Sequencing and Transcriptomic Analyses
RNA-sequencing was performed by NovoGene as previ-
ously described by us (54). Standard 150-bp paired-end
sequencing was performed, and read counts were >20 mil-
lion. Raw counts from RNA-sequencing were used as inputs
into Partek Flow. Alignment was performed with STAR with
mmu39. After low-expressed genes were ltered, DESeq2
(version 1.34.0) was used for normalization and differential
analyses to identify differentially expressed genes (DEGs)
with pairwise comparisons (56). DEGs were identied with a
false discovery rate (FDR, step-up procedure) adjusted P
value <0.05. DEGs with adjusted Pvalue <0.05 were used
for downstream functional analysis using ConsensusPath DB
(57). Pathway analysis was conducted using the mouse over-
representation feature, up- or downregulated DEGs, and the
Reactome database with default settings. For pathway analy-
sis, qvalues were used to determine signicance based on
DEGs with adjusted P<0.05.
Proteomics: FASP bHPLC-Orbitrap Fusion
Proteomics was performed at the University of Arkansas
for Medical Sciences Proteomics Core. Protein samples were
reduced, alkylated, and digested by lter-aided sample prep-
aration (58) with sequencing-grade modied porcine trypsin
(Promega). Tryptic peptides were separated into 46 fractions
on a 100 1.0-mm Acquity BEH C18 column (Waters) with
an UltiMate 3000 UHPLC system (Thermo) with a 50-min
gradient from 99:1 to 60:40 buffer A-to-Bratio under basic
pH conditions and then consolidated into 12 superfractions;
buffer A = 0.1% formic acid, 0.5% acetonitrile, and buffer B =
0.1% formic acid, 99.9% acetonitrile. Each superfraction was
then further separated by reverse-phase XSelect CSH C18 2.5-
μm resin (Waters) on an in-line 150 0.075-mm column
with an UltiMate 3000 RSLCnano system (Thermo). Peptides
were eluted with a 45-min gradient from 98:2 to 65:35 buffer
A-to-Bratio. Eluted peptides were ionized by electrospray
(2.4 kV) followed by mass spectrometric analysis on an
Orbitrap Fusion Tribrid mass spectrometer (Thermo). MS
data were acquired with the FTMS analyzer in prole mode
at a resolution of 240,000 over a range of 375 to 1,500 m/z.
After HCD activation, MS/MS data were acquired with the
ion trap analyzer in centroid mode and normal mass range
with normalized collision energy of 2831% depending on
charge state and precursor selection range. Proteins were
identied by database search using MaxQuant (Max Planck
Institute) label-free quantication with a parent ion toler-
ance of 2.5 ppm and a fragment ion tolerance of 0.5 Da.
Scaffold Q þS (Proteome Software) was used to verify MS/
MS-based peptide and protein identications. Protein identi-
cations were accepted if they could be established with
<1.0% false discovery and contained at least two identied
peptides. Protein probabilities were assigned by the Protein
Prophet algorithm (59). Protein abundance is presented with
intensity-based absolute quantication (iBAQ) values. To
determine differential expression and pathway analysis
between conditions, Pvalues based on unpaired nondirec-
tional ttests were employed with a cutoff of P<0.05.
BenjaminiHochberg-adjusted Pvalues were subsequently
calculated from these Pvalues. Untranslated region (UTR)
sequences for genes that resulted in proteins of interest were
downloaded from UCSC Genome Browser and used for pre-
diction of miR-16 binding sites. The potential binding site
was identied with RNAhybrid software (60,61), which com-
bines thermodynamic and seed sequence information.
RESULTS
miR-16 Is Lower after 72-h Synergist Ablation
Mechanical Overload of the Plantaris Muscle
Previous work suggests that muscle miR-16 levels are
inversely related to postnatal muscle growth in chickens
(34,36) and are lower during regeneration in mice (24). We
showed that miR-16 is lower in muscle tissue after a bout of
resistance exercise in rats (37). Previous work showed signi-
cant changes in global muscle miRNA levels in the early
phase of mechanical overload-induced hypertrophy (62). To
determine whether miR-16 declines during the early phase
of a well-characterized model of rapid loading-induced hy-
pertrophyinmice(49,63,64), we compared miR-16-5p levels
in 72-h mechanically overloaded (MOV) plantaris muscle to
sham-operated muscles. miR-16 was 28% lower in muscle tis-
sue during MOV (P=0.06)(Fig. 1A). Overall, low miR-16
appears characteristic of growing muscle in several settings
and species (6,3537).
miR-16 Inuences Ribosomal-, p53-, and Pol II
Transcription-Related mRNA Abundance
The in vitro study design is found in Fig. 1B.Thetransfec-
tion of a miR-16 inhibitor was effective at knocking down
miR-16 (miR-16 KD) in proliferating C2C12 myoblasts (Fig. 1C).
Inhibition of miR-16 in myoblasts upregulated mRNA levels
of Rps12,Rps26,Rplp1,Rpl22l1,andRpl35 (adj. P<0.05),
which encode ribosomal proteins (Fig. 1D). Additional ribo-
somal protein genes tended to be higher after miR-16 knock-
down (Rpl27 and Rpl27a,adj.P<0.10) (Supplemental Table
S1). The major processes downregulated by miR-16 inhibition
miR-16 IN MYOGENIC CELLS
AJP-Cell Physiol doi:10.1152/ajpcell.00071.2023 www.ajpcell.org C1103
Downloaded from journals.physiology.org/journal/ajpcell at Univ of Arkansas Fayetteville (130.184.252.076) on May 18, 2023.
were Regulation of TP53 Activity through Methylation
(Reactome, q= 0.0014) and RNA Polymerase II Transcription
(Reactome, q= 0.0026) (Supplemental Table S2). In the for-
mer pathway, Atm,Ep300,andMdm2 were lower (adj. P<
0.05). In addition to those three genes, Aff4,Crebbp,Kmt2a,
Kmt2c, Mga,Nr2c2,andSox9 were lower after miR-16 knock-
down in the latter pathway (adj. P<0.05) (Fig. 1E). The long
noncoding RNA (lncRNA) H19 was elevated with miR-16 inhi-
bition (adj. P= 0.0035), and the lncRNAs Malat1 (adj. P=
0.059) and Xist (adj. P= 0.009) were lower (Supplemental
Table S1).
miR-16 Knockdown Increases Metabolism-Related
Proteins and Specic Markers of Myogenic Cell
Dierentiation
Proteomics was performed in ve technical replicates per
condition. After miR-16 knockdown, ve proteins were upregu-
lated (EEF1A2, HIST1H1B, NPC2, PRPH, and TSEN54) and
seven proteins were downregulated (GLRX2, GOLPH3, GRK6,
LAPTM4A, NDC1, THAP4, and ZC3H11A) with an adjusted
Pvalue <0.05, and 899 proteins achieved signicance at P<
0.05 (Supplemental Table S3). Proteins involved in different
aspects of Metabolism (Reactome, q= 0.0178) were upregulated
(assessed using proteins with P<0.05) (Fig. 2A)(Supplemental
Tables S3 and S4). Within the broad Metabolism pathway,
various tricarboxylic acid (TCA) cycle proteins were selec-
tively enriched (Reactome, q= 0.00293). Proteins related
to Metabolism of RNA (Reactome, q= 0.000456) were
most downregulated by miR-16 inhibition (Supplemental
Table S4). Markers of myogenic cell maturation, including
myosin light chains (MYL6B, MYL9, MYL12A, P<0.05) as
well as smooth muscle actin (ACTA2, P=0.04,adj.P=
0.41) (Fig. 2B), were upregulated with miR-16 KD. Specic
proteins that have a dened role in satellite cell behavior
were also altered. In addition to muscle-enriched EEF1A2
(P= 0.0000164, adj. P=0.018)(Fig. 2C)(6567), OPA1 (P<
0.05, adj. P=0.25and0.30)(Fig. 2D)(68,69)andPTEN
(P<0.05, adj. P=0.41)(Fig. 2E)(7072)wereelevatedby
miR-16 inhibition; the latter two proteins are of interest
butdidnotachievesignicance according to adjusted
Aff4 Atm Crebbp Ep300 Kmt2a Kmt2c Mdm2 Mga Nr2c2 Sox9
0
1000
2000
3000
4000
5000
DESeq2 Gene Expression
Regulation of p53 and Pol II Transcription mRNA Levels
Control
miR-16 KD
0.0
0.5
1.0
1.5
miR-16
*
*
*
**
*
*
*
*
miR-16 : U6
Rps12 Rps26 Rplp1 Rpl22l1 Rpl35
0
5000
10000
15000
DESeq2 Gene Expression
Ribosomal Protein mRNA Levels
Control
miR-16 KD
Sham MOV
0.0
0.5
1.0
1.5 miR-16
p=0.06
A
D
E
miR-16 : U6
BC
*
*
*
**
*
*
Figure 1. miR-16 levels in overloaded mus-
cle in vivo and RNA-sequencing (RNA-seq)
analysis in proliferating C2C12 myoblasts 18
h after miR-16 knockdown (KD). A:miR-16
levels in plantaris skeletal muscle after 72 h
of synergist ablation mechanical overload
of the plantaris (MOV) relative to sham oper-
ation. B: in vitro study design. C:miR-16lev-
els in control and KD myoblasts. D:levelsof
ribosomal protein genes in control and KD
myoblasts. E:levelsofp53-andPolIItran-
scription-related genes in control and KD
myoblasts. For RNA-seq experiments, n=3
control and n= 2 KD technical replicates.
adj. P<0.05.
miR-16 IN MYOGENIC CELLS
C1104 AJP-Cell Physiol doi:10.1152/ajpcell.00071.2023 www.ajpcell.org
Downloaded from journals.physiology.org/journal/ajpcell at Univ of Arkansas Fayetteville (130.184.252.076) on May 18, 2023.
Pvalues. OPA1 was detected twice (canonical 111 kDa and
truncated 34 kDa), and the levels of both isoforms were
higher with knockdown (P<0.05) but not according to
adjusted Pvalues.
An overlap in gene and protein expression would not neces-
sarilybeexpectedatadiscretetimepointsinceproteintransla-
tion lags behind changes in mRNA expression. Nevertheless,
comparison of gene to protein levels revealed agreement for
Bmpr2,Cltc,Lama5,Lpp,Mga,Nptx1,andSema5a,allofwhich
were lowered by miR-16 knockdown (P<0.05). Rpl27a,which
was trending to increase at the gene level, was elevated at the
protein level (P=0.01,adj.P=0.28).Sinceweobservedupregu-
lation of proteins with miR-16 knockdown, and miRNAs can
prevent protein translation without altering transcript levels
Normalized iBAQ
PTEN
110 kDa 34 kDa
Normalized iBAQ
OPA1
*
*
0.0
5.0×109
1.0×1010
1.5×1010
0.0
5.0×106
1.0×107
1.5×107
2.0×107
Normalized iBAQ
ACTA2
Control
miR-16 KD
A
-3 -2 -1 0
Dual Incision in GG-NER
Chromatin modifying enzymes
Chromatin organization
Oocyte meiosis - Mus musculus (mouse)
Death Receptor Signalling
DNA Replication
COPI-dependent Golgi-to-ER retrograde traffic
Toll-like Receptor Cascades
Nucleotide Excision Repair
SUMOylation of DNA damage response and repair proteins
Transport of Mature mRNA Derived from an Intronless Transcript
Resolution of Abasic Sites (AP sites)
Resolution of AP sites via the multiple-nucleotide patch replacement pathway
Free fatty acids regulate insulin secretion
Interleukin-2 signaling
Synthesis of DNA
MAP kinase activation
Interleukin-17 signaling
S Phase
Toll Like Receptor 3 (TLR3) Cascade
TRAF6 mediated induction of NFkB and MAP kinases upon TLR7/8 or 9 activation
MyD88 cascade initiated on plasma membrane
Toll Like Receptor 10 (TLR10) Cascade
Toll Like Receptor 5 (TLR5) Cascade
Mitochondrial translation termination
MyD88 dependent cascade initiated on endosome
MyD88:MAL(TIRAP) cascade initiated on plasma membrane
Toll Like Receptor TLR1:TLR2 Cascade
Toll Like Receptor TLR6:TLR2 Cascade
Toll Like Receptor 2 (TLR2) Cascade
Mitochondrial translation
Toll Like Receptor 7/8 (TLR7/8) Cascade
Activation of the AP-1 family of transcription factors
Inactivation of APC/C via direct inhibition of the APC/C complex
Inhibition of the proteolytic activity of APC/C
Cell Cycle, Mitotic
Golgi-to-ER retrograde transport
Toll Like Receptor 9 (TLR9) Cascade
MAPK targets/ Nuclear events mediated by MAP kinases
Metabolism
Citric acid cycle (TCA cycle)
Pyruvate metabolism and Citric Acid (TCA) cycle
Log10 q value
UPregulated Proteins: Reactome
Normalized iBAQ
EEF1A2
0.0
1.0×109
2.0×109
3.0×109
4.0×109
5.0×109
0.0
1.0×106
2.0×106
3.0×106
4.0×106
5.0×106
p=0.04
adj.
p=0.41
p=0.005
adj.
p=0.25
p=0.013
adj.
p=0.30 p=0.04
adj.
p=0.41
p<0.001
adj.
p=0.018
*
BC
DE
**
Figure 2. Proteomic analysis in proliferating
C2C12 myoblasts 18 h after miR-16 knock-
down (KD). A: Reactome pathway analysis
using upregulated differentially expressed
proteins (P<0.05). BE: thermodynamic and
seed sequence target prediction for miR-16
and protein levels of ACTA2 (B), EEF1A2 (C),
OPA1 (D), and PTEN (E) in control and KD
myoblasts. iBAQ, intensity-based absolute
quantication. For all experiments, n= 5 tech-
nical replicates were utilized. P<0.05.
Created with BioRender.com.
miR-16 IN MYOGENIC CELLS
AJP-Cell Physiol doi:10.1152/ajpcell.00071.2023 www.ajpcell.org C1105
Downloaded from journals.physiology.org/journal/ajpcell at Univ of Arkansas Fayetteville (130.184.252.076) on May 18, 2023.
(14,73), we performed a 30UTR binding afnity analysis of
miR-16 for ACTA2, EEF1A2, OPA1, and PTEN (Fig. 2, BE),
using RNAhybrid (60,61).Basedonfreeenergy,theresultscol-
lectively suggested that miR-16 could regulate the levels of
these proteins. The highest complementarity between the miR-
16 seed sequence and 30UTR sequence was for EEF1A2, OPA1,
and PTEN (Fig. 2. CE).
DISCUSSION
Satellite cells differentiate and begin fusing appreciably to
muscle bers between days 4 and 5 of mechanical overload in
adult mice (25)and5 days after transplantation into muscle
of mice (74). In our 72-h overload experiments, the contribu-
tions of miR-16 from satellite cells versus other cell types in
muscle tissue (e.g., muscle bers or immune cells) (75) cannot
be discerned. The inuence of hypertrophy per se versus a
degeneration/regeneration response that can occur with syner-
gist ablation (19,76) is also unclear. Nevertheless, our observa-
tion of lower miR-16 in murine muscle tissue after 72 h of MOV
corresponds with reduced muscle miR-16 levels during early
in vivo regeneration after injury (6) and during recovery from
a bout of resistance exercise (37). To model miR-16 regulation
of myogenic cell behavior, we repressed it in proliferating
myoblasts and performed transcriptomic and proteomic prol-
ing.ThesedataprovideaframeworkforunderstandingmiR-
16s role in myogenic cell fate.
miR-16-5p can induce p53 signaling in myogenic cells (34).
Repression of p53-related gene expression with miR-16
knockdown dovetails with this observation. In C2C12s in
vitro, the expression of ribosomal proteins is upregulated
during early differentiation (77,78). Enrichment of ribo-
somal protein genes is associated with satellite cell fusion
into the myober syncytium during hypertrophy (79). The
induction of ribosomal protein genes, specically Rps26,by
miR-16 inhibition may characterize myogenic cells that are
primed for fusion (79). The lncRNA H19 was elevated with
miR-16 knockdown, whereas Malat1 and Xist were lower.
lncRNAs are not typically targeted for decay by miRNAs
since translation is seemingly required for RNA destabiliza-
tion (80). H19 induction (8184)andMalat1 reduction (85)
are strongly implicated in myogenic cell differentiation, but
how miR-16 could affect lncRNA levels in myogenic cells
deserves further investigation. In concert with evidence sug-
gesting that miR-16 can target Myomaker, the gatekeeper of
myogenic cell fusion that gradually increases during differ-
entiation (86,87), our RNA-seq data collectively point to
declining miR-16 levels facilitating lineage progression to-
ward differentiation.
At the protein level, repressing miR-16 in proliferating myo-
blasts leads to several alterations that are indicative of miR-
16s roles in the regulation of differentiation. The enrichment
of myosin light chains, noted in our data, is a sign of myo-
genic cell maturation (88,89). Upregulation of smooth muscle
actin (Acta2) is also strongly associated with myoblast differ-
entiation (9093). Satellite cell differentiation is controlled by
a metabolic shift and progression from glycolysis to the TCA
cycle (94). Elevated TCA cycle proteins with miR-16 repres-
sion therefore seems intuitive from a metabolic perspective.
Induction and phosphorylation of EEF1A2 is linked to myo-
genic cell differentiation (6567). EEF1A2 is muscle enriched
(95) and highly regulated during muscle development (96),
protects myotubes from cell death (65), and controls Utrophin
levels in skeletal muscle (97,98). EEF1A2 is also a core compo-
nent of cardiomyocyte differentiation (99). The causal func-
tions of EEF1A2 control by miR-16 in myoblast differentiation
deserve further study. Higher PTEN and OPA1 with miR-16 in-
hibition is noteworthy since both are implicated in reinforc-
ing satellite cell quiescence (68,70,71). PTEN can repress the
satellite cell identity gene Pax7 (72,100)butmayalsofacili-
tate a return to quiescence (70,71) that could ensue if differen-
tiation and fusion does not progress. Whereas OPA1 supports
quiescence by maintaining mitochondrial integrity (68), OPA1
induction and mitophagy is an essential component of suc-
cessful C2C12 myoblast differentiation (69).
The inhibition of miR-16 in proliferating myoblasts, which
occurs naturally during myogenic differentiation, reveals its
contributionstothisprocess.Specically, we uncover miR-
16s regulation of ribosomal, p53-related, and lncRNA gene
expression as well as metabolic- and muscle maturation-
related protein abundance. A more detailed understanding of
miR-16 dynamics, specically in satellite cells in vivo during
different stages of myogenesis, regeneration, and hypertrophy,
will inform how miR-16 controls muscle mass in varying cir-
cumstances. The present study is limited to a single time point
with an immortalized cell line. We also do not report on myo-
genic cell behavior after miR-16 inhibition, although this has
been documented in detail elsewhere (6,34,40). Limitations
aside, we provide insights from two -omic layers to expand our
understanding of miR-16 regulation in myogenic cells.
DATA AVAILABILITY
Processed data are provided in Supplemental Tables S1S4
and were deposited in GEO (GSE229134).
SUPPLEMENTAL DATA
Supplemental Tables S1S4: https://doi.org/10.6084/m9.gshare.
22151624.
ACKNOWLEDGMENTS
The authors thank Dr. John J. McCarthy of the University of
Kentucky and Dr. Vandre C. Figueiredo of Oakland University for
critical feedback on this manuscript. The authors also thank the
University of Arkansas for Medical Sciences Proteomics Core for
conducting the proteomics and analysis.
The Graphical Abstract was generated with BioRender.
GRANTS
This work was supported by NIH Grant R00 AG063994 to K.A.M.
Support for proteomics was provided by an Arkansas Biosciences
Institute grant to N.P.G.
DISCLOSURES
No conicts of interest, nancial or otherwise, are declared by
the authors.
AUTHOR CONTRIBUTIONS
N.P.G. and K.A.M. conceived and designed research; S.L., D.L.,
F.M., P.J.K., and F.v. performed experiments; S.L., F.M., P.J.K., I.J.V.,
miR-16 IN MYOGENIC CELLS
C1106 AJP-Cell Physiol doi:10.1152/ajpcell.00071.2023 www.ajpcell.org
Downloaded from journals.physiology.org/journal/ajpcell at Univ of Arkansas Fayetteville (130.184.252.076) on May 18, 2023.
and K.A.M. analyzed data; S.L., F.M., I.J.V., and K.A.M. interpreted
results of experiments; S.L., I.J.V., and K.A.M. prepared gures; S.L.,
D.L., and K.A.M. drafted manuscript; S.L., D.L., F.M., P.J.K., I.J.V., F.v.,
N.P.G., and K.A.M. edited and revised manuscript; S.L., D.L., F.M.,
P.J.K., I.J.V., F.v., N.P.G., and K.A.M. approved nal version of
manuscript.
REFERENCES
1. Castel D,Baghdadi MB,Mella S,Gayraud-Morel B,Marty V,
Cavaill
eJ,Antoniewski C,Tajbakhsh S. Small-RNA sequencing
identies dynamic microRNA deregulation during skeletal muscle
lineage progression. Sci Rep 8: 4208, 2018. doi:10.1038/s41598-018-
21991-w.
2. Cheung TH,Quach NL,Charville GW,Liu L,Park L,Edalati A,Yoo
B,Hoang P,Rando TA. Maintenance of muscle stem-cell quies-
cence by microRNA-489. Nature 482: 524528, 2012. doi:10.1038/
nature10834.
3. Crist CG,Montarras D,Buckingham M. Muscle satellite cells are
primed for myogenesis but maintain quiescence with sequestration
of Myf5 mRNA targeted by microRNA-31 in mRNP granules. Cell
Stem Cell 11: 118126, 2012. [Erratum in Cell Stem Cell 11: 279, 2012].
doi:10.1016/j.stem.2012.03.011.
4. Koning M,Werker PM,van Luyn MJ,Krenning G,Harmsen MC. A
global downregulation of microRNAs occurs in human quiescent sat-
ellite cells during myogenesis. Differentiation 84: 314321, 2012.
doi:10.1016/j.diff.2012.08.002.
5. de Morree A,Klein JD,Gan Q,Farup J,Urtasun A,Kanugovi A,
Bilen B,van Velthoven CT,Quarta M,Rando TA. Alternative poly-
adenylation of Pax3 controls muscle stem cell fate and muscle func-
tion. Science 366: 734738, 2019. doi:10.1126/science.aax1694.
6. Farina NH,Hausburg M,Betta N,Pulliam C,Srivastava D,
Cornelison D,Olwin BB. AroleforRNApost-transcriptionalregula-
tion in satellite cell activation. Skelet Muscle 2: 21, 2012. doi:10.1186/
2044-5040-2-21.
7. Boutet SC,Cheung TH,Quach NL,Liu L,Prescott SL,Edalati A,Iori
K,Rando TA. Alternative polyadenylation mediates microRNA regu-
lation of muscle stem cell function. Cell Stem Cell 10: 327336, 2012.
doi:10.1016/j.stem.2012.01.017.
8. Chen JF,Tao Y,Li J,Deng Z,Yan Z,Xiao X,Wang D-Z. microRNA-1
and microRNA-206 regulate skeletal muscle satellite cell prolifera-
tion and differentiation by repressing Pax7. JCellBiol190: 867879,
2010. doi:10.1083/jcb.200911036.
9. Lozano-Velasco E,Vallejo D,Esteban FJ,Doherty C,Hernández-
Torres F,Franco D,Aránega AE. A Pitx2-MicroRNA pathway modu-
lates cell proliferation in myoblasts and skeletal-muscle satellite cells
and promotes their commitment to a myogenic cell fate. Mol Cell
Biol 35: 28922909, 2015. doi:10.1128/MCB.00536-15.
10. Su Y,Yu Y,Liu C,Zhang Y,Liu C,Ge M,Li L,Lan M,Wang T,Li M,
Liu F,Xiong L,Wang K,He T,Shi J,Song Y,Zhao Y,Li N,Yu Z,
Meng Q. Fate decision of satellite cell differentiation and self-
renewal by miR-31-IL34 axis. Cell Death Differ 27: 949965, 2020.
doi:10.1038/s41418-019-0390-x.
11. Baghdadi MB,Firmino J,Soni K,Evano B,Di Girolamo D,Mourikis
P,Castel D,Tajbakhsh S. Notch-induced miR-708 antagonizes sat-
ellite cell migration and maintains quiescence. Cell Stem Cell 23:
859868.e5, 2018. doi:10.1016/j.stem.2018.09.017.
12. Goljanek-Whysall K,Sweetman D,Abu-Elmagd M,Chapnik E,
Dalmay T,Hornstein E,M
unsterberg A. MicroRNA regulation of the
paired-box transcription factor Pax3 confers robustness to develop-
mental timing of myogenesis. Proc Natl Acad Sci USA 108: 11936
11941, 2011. doi:10.1073/pnas.1105362108.
13. Goljanek-Whysall K,Pais H,Rathjen T,Sweetman D,Dalmay T,
M
unsterberg A. Regulation of multiple target genes by miR-1 and
miR-206 is pivotal for C2C12 myoblast differentiation. JCellSci125:
35903600, 2012. doi:10.1242/jcs.101758.
14. Fry CS,Kirby TJ,Kosmac K,McCarthy JJ,Peterson CA. Myogenic
progenitor cells control extracellular matrix production by broblasts
during skeletal muscle hypertrophy. Cell Stem Cell 20: 5669, 2017.
doi:10.1016/j.stem.2016.09.010.
15. Bachman JF,Klose A,Liu W,Paris ND,Blanc RS,Schmalz M,
Knapp E,Chakkalakal JV. Prepubertal skeletal muscle growth
requires Pax7-expressing satellite cell-derived myonuclear contribu-
tion. Development 145: dev167197, 2018. doi:10.1242/dev.167197.
16. Cramer AA,Prasad V,Eftestøl E,Song T,Hansson KA,Dugdale
HF,Sadayappan S,Ochala J,Gundersen K,Millay DP. Nuclear
numbers in syncytial muscle bers promote size but limit the devel-
opment of larger myonuclear domains. Nat Commun 11: 6287, 2020.
doi:10.1038/s41467-020-20058-7.
17. Hansson KA,Eftestøl E,Bruusgaard JC,Juvkam I,Cramer AW,
Malthe-Sørenssen A,Millay DP,Gundersen K. Myonuclear content
regulates cell size with similar scaling properties in mice and
humans. Nat Commun 11: 6288, 2020. doi:10.1038/s41467-020-
20057-8.
18. Fry CS,Lee JD,Mula J,Kirby TJ,Jackson JR,Liu F,Yang L,
Mendias CL,Dupont-Versteegden EE,McCarthy JJ,Peterson CA.
Inducible depletion of satellite cells in adult, sedentary mice impairs
muscle regenerative capacity without affecting sarcopenia. Nat Med
21: 7680, 2015. doi:10.1038/nm.3710.
19. McCarthy JJ,Mula J,Miyazaki M,Erfani R,Garrison K,Farooqui
AB,Srikuea R,Lawson BA,Grimes B,Keller C,Van Zant G,
Campbell KS,Esser KA,Dupont-Versteegden EE,Peterson CA.
Effective ber hypertrophy in satellite cell-depleted skeletal muscle.
Development 138: 36573666, 2011. doi:10.1242/dev.068858.
20. Lepper C,Partridge TA,Fan CM. An absolute requirement for Pax7-
positive satellite cells in acute injury-induced skeletal muscle regener-
ation. Development 138: 36393646, 2011. doi:10.1242/dev.067595.
21. Sambasivan R,Yao R,Kissenpfennig A,Van Wittenberghe L,Paldi
A,Gayraud-Morel B,Guenou H,Malissen B,Tajbakhsh S,Galy A.
Pax7-expressing satellite cells are indispensable for adult skeletal
muscle regeneration. Development 138: 36473656, 2011. [Erratum
in Development 138: 4333, 2011]. doi:10.1242/dev.067587.
22. Murphy MM,Lawson JA,Mathew SJ,Hutcheson DA,Kardon G.
Satellite cells, connective tissue broblasts and their interactions are
crucial for muscle regeneration. Development 138: 36253637, 2011.
doi:10.1242/dev.064162.
23. Englund D,Figueiredo V,Dungan C,Murach K,Peck B,Petrosino
J,Brightwell C,Dupont A,Neal A,Fry C,Accornero F,McCarthy J,
Peterson C. Satellite cell depletion disrupts transcriptional coordina-
tion and muscle adaptation to exercise. Function (Oxf) 2: zqaa033,
2021. doi:10.1093/function/zqaa033.
24. Englund DA,Murach KA,Dungan CM,Figueiredo VC,Vechetti IJ Jr,
Dupont-Versteegden EE,McCarthy JJ,Peterson CA. Depletion of
resident muscle stem cells negatively impacts running volume, physi-
cal function and muscle hypertrophy in response to lifelong physical
activity. Am J Physiol Cell Physiol 318: C1178C1188, 2020. doi:10.1152/
ajpcell.00090.2020.
25. Murach KA,Peck BD,Policastro RA,Vechetti IJ,Van Pelt DW,
Dungan CM,Denes LT,Fu X,Brightwell CR,Zentner GE,Dupont-
Versteegden EE,Richards CI,Smith JJ,Fry CS,McCarthy JJ,
Peterson CA. Early satellite cell communication creates a permissive
environment for long-term muscle growth. iScience 24: 102372,
2021. doi:10.1016/j.isci.2021.102372.
26. Murach KA,Vechetti IJ Jr,Van Pelt DW,Crow SE,Dungan CM,
Figueiredo VC,Kosmac K,Fu X,Richards CI,Fry CS,McCarthy JJ,
Peterson CA. Fusion-independent satellite cell communication to mus-
cle bers during load-induced hypertrophy. Function (Oxf) 1: zqaa009,
2020. doi:10.1093/function/zqaa009.
27. Murach KA,Fry CS,Dupont-Versteegden EE,McCarthy JJ,
Peterson CA. Fusion and beyond: Satellite cell contributions to load-
ing-induced skeletal muscle adaptation. FASEB J 35: e21893, 2021.
doi:10.1096/fj.202101096R.
28. Murach KA,Fry CS,Kirby TJ,Jackson JR,Lee JD,White SH,Dupont-
Versteegden EE,McCarthy JJ,Peterson CA. Starring or supporting
role? Satellite cells and skeletal muscle ber size regulation. Physiology
(Bethesda) 33: 2638, 2018. doi:10.1152/physiol.00019.2017.
29. Bagley JR,Denes LT,McCarthy JJ,Wang ET,Murach KA. The myo-
nuclear domain in adult skeletal muscle bres: past, present, and
future. JPhysiol601: 723741, 2023. doi:10.1113/JP283658.
30. Yue J,Tigyi G. Conservation of miR-15a/16-1 and miR-15b/16-2 clus-
ters. Mamm Genome 21: 8894, 2010. doi:10.1007/s00335-009-
9240-3.
31. Liang H,Fu Z,Jiang X,Wang N,Wang F,Wang X,Zhang S,Wang
Y,Yan X,Guan WX,Zhang CY,Zen K,Zhang Y,Chen X,Zhou G.
miR-16 promotes the apoptosis of human cancer cells by targeting
FEAT. BMC Cancer 15: 448, 2015. doi:10.1186/s12885-015-1458-8.
miR-16 IN MYOGENIC CELLS
AJP-Cell Physiol doi:10.1152/ajpcell.00071.2023 www.ajpcell.org C1107
Downloaded from journals.physiology.org/journal/ajpcell at Univ of Arkansas Fayetteville (130.184.252.076) on May 18, 2023.
32. Aqeilan R,Calin G,Croce C. miR-15a and miR-16-1 in cancer: discov-
ery, function and future perspectives. Cell Death Differ 17: 215220,
2010. doi:10.1038/cdd.2009.69.
33. Calin GA,Dumitru CD,Shimizu M,Bichi R,Zupo S,Noch E,Aldler H,
Rattan S,Keating M,Rai K,Rassenti L,Kipps T,Negrini M,Bullrich
F,Croce CM. Frequent deletions and down-regulation of micro-RNA
genes miR15 and miR16 at 13q14 in chronic lymphocytic leukemia.
Proc Natl Acad Sci USA 99: 1552415529, 2002. doi:10.1073/
pnas.242606799.
34. Cai B,Ma M,Chen B,Li Z,Abdalla BA,Nie Q,Zhang X. MiR-16-5p
targets SESN1 to regulate the p53 signaling pathway, affecting myo-
blast proliferation and apoptosis, and is involved in myoblast differen-
tiation. Cell Death Dis 9: 367, 2018. doi:10.1038/s41419-018-0403-6.
35. Jia X,Lin H,Nie Q,Zhang X,Lamont SJ. A short insertion mutation dis-
rupts genesis of miR-16 and causes increased body weight in domesti-
cated chicken. Sci Rep 6: 36433, 2016. doi:10.1038/srep36433.
36. Jia X,Ouyang H,Abdalla BA,Xu H,Nie Q,Zhang X. miR-16 controls
myoblast proliferation and apoptosis through directly suppressing
Bcl2 and FOXO1 activities. Biochim Biophys Acta Gene Regul Mech
1860: 674684, 2017. doi:10.1016/j.bbagrm.2017.02.010.
37. Lee DE,Brown JL,Rosa ME,Brown LA,Perry RA Jr,Wiggs MP,
Nilsson MI,Crouse SF,Fluckey JD,Washington TA,Greene NP.
microRNA-16 is downregulated during insulin resistance and con-
trols skeletal muscle protein accretion. JCellBiochem117: 1775
1787, 2016. doi:10.1002/jcb.25476.
38. Aguilar CA,Pop R,Shcherbina A,Watts A,Matheny RW Jr,
Cacchiarelli D,Han WM,Shin E,Nakhai SA,Jang YC,Carrigan CT,
Gifford CA,Kottke MA,Cesana M,Lee J,Urso ML,Meissner A.
Transcriptional and chromatin dynamics of muscle regeneration af-
ter severe trauma. Stem Cell Reports 7: 983997, 2016. doi:10.1016/j.
stemcr.2016.09.009.
39. Siengdee P,Trakooljul N,Murani E,Schwerin M,Wimmers K,
Ponsuksili S. MicroRNAs regulate cellular ATP levels by targeting mito-
chondrial energy metabolism genes during C2C12 myoblast differentia-
tion. PLoS One 10: e0127850, 2015. doi:10.1371/journal.pone.0127850.
40. Harding RL,Velleman SG. MicroRNA regulation of myogenic satel-
lite cell proliferation and differentiation. Mol Cell Biochem 412: 181
195, 2016. doi:10.1007/s11010-015-2625-6.
41. Cornelison D,Olwin BB,Rudnicki MA,Wold BJ. MyoD
/
satellite
cells in single-ber culture are differentiation defective and MRF4 de-
cient. Dev Biol 224: 122137, 2000. doi:10.1006/dbio.2000.9682.
42. Sabourin LA,Girgis-Gabardo A,Seale P,Asakura A,Rudnicki MA.
Reduced differentiation potential of primary MyoD
/
myogenic
cells derived from adult skeletal muscle. JCellBiol144: 631643,
1999. doi:10.1083/jcb.144.4.631.
43. Ishibashi J,Perry RL,Asakura A,Rudnicki MA. MyoD induces myo-
genic differentiation through cooperation of its NH
2
-and COOH-termi-
nal regions. JCellBiol171: 471482, 2005. doi:10.1083/jcb.200502101.
44. Penn BH,Bergstrom DA,Dilworth FJ,Bengal E,Tapscott SJ. A
MyoD-generated feed-forward circuit temporally patterns gene
expression during skeletal muscle differentiation. Genes Dev 18:
23482353, 2004. doi:10.1101/gad.1234304.
45. Yablonka-Reuveni Z,Rudnicki MA,Rivera AJ,Primig M,Anderson
JE,Natanson P. The transition from proliferation to differentiation is
delayed in satellite cells from mice lacking MyoD. Dev Biol 210:
440455, 1999. doi:10.1006/dbio.1999.9284.
46. Di Carlo A,De Mori R,Martelli F,Pompilio G,Capogrossi MC,
Germani A. Hypoxia inhibits myogenic differentiation through accel-
erated MyoD degradation. JBiolChem279: 1633216338, 2004.
doi:10.1074/jbc.M313931200.
47. Zhang H,Wen J,Bigot A,Chen J,Shang R,Mouly V,Bi P. Human
myotube formation is determined by MyoDMyomixer/Myomaker
axis. Sci Adv 6: eabc4062, 2020. doi:10.1126/sciadv.abc4062.
48. Yamamoto M,Legendre NP,Biswas AA,Lawton A,Yamamoto S,
Tajbakhsh S,Kardon G,Goldhamer DJ. Loss of MyoD and Myf5 in
skeletal muscle stem cells results in altered myogenic programming
and failed regeneration. Stem Cell Reports 10: 956969, 2018.
doi:10.1016/j.stemcr.2018.01.027.
49. Kirby TJ,McCarthy JJ,Peterson CA,Fry CS. Synergist ablation as a
rodent model to study satellite cell dynamics in adult skeletal mus-
cle. Methods Mol Biol 1460: 4352, 2016. doi:10.1007/978-1-4939-
3810-0_4.
50. Seale P,Sabourin LA,Girgis-Gabardo A,Mansouri A,Gruss P,
Rudnicki MA. Pax7 is required for the specication of myogenic
satellite cells. Cell 102: 777786, 2000. doi:10.1016/s0092-8674(00)
00066-0.
51. Asakura A,Rudnicki MA,Komaki M. Muscle satellite cells are multi-
potential stem cells that exhibit myogenic, osteogenic, and adipo-
genic differentiation. Differentiation 68: 245253, 2001. doi:10.1046/
j.1432-0436.2001.680412.x.
52. Lamon S,Zacharewicz E,Stephens AN,Russell AP. EPO-receptor
is present in mouse C2C12 and human primary skeletal muscle cells
but EPO does not inuence myogenesis. Physiol Rep 2: e00256,
2014. doi:10.1002/phy2.256.
53. Yaffe D,Saxel O. Serial passaging and differentiation of myogenic
cells isolated from dystrophic mouse muscle. Nature 270: 725727,
1977. doi:10.1038/270725a0.
54. Murach KA,Liu Z,Jude B,Figueiredo VC,Wen Y,Khadgi S,Lim S,
Morena da Silva F,Greene NP,Lanner JT,McCarthy JJ,Vechetti IJ
Jr,von Walden F. Multi-transcriptome analysis following an acute
skeletal muscle growth stimulus yields tools for discerning global
and MYC regulatory networks. JBiolChem298: 102515, 2022.
doi:10.1016/j.jbc.2022.102515.
55. Lim S,Deaver JW,Rosa-Caldwell ME,Haynie WS,Morena da Silva
F,Cabrera AR,Schrems ER,Saling LW,Jansen LT,Dunlap KR,
Wiggs MP,Washington TA,Greene NP. Development of metabolic
and contractile alterations in development of cancer cachexia in
female tumor-bearing mice. J Appl Physiol (1985) 132: 5872, 2022.
doi:10.1152/japplphysiol.00660.2021.
56. Love MI,Huber W,Anders S. Moderated estimation of fold change
and dispersion for RNA-seq data with DESeq2. Genome Biol 15:
550, 2014. doi:10.1186/s13059-014-0550-8.
57. Kamburov A,Herwig R. ConsensusPathDB 2022: molecular interac-
tions update as a resource for network biology. Nucleic Acids Res
50: D587D595, 2022. doi:10.1093/nar/gkab1128.
58. Wis¨niewski JR,Zougman A,Nagaraj N,Mann M. Universal sample
preparation method for proteome analysis. Nat Methods 6: 359
362, 2009. doi:10.1038/nmeth.1322.
59. Nesvizhskii AI,Keller A,Kolker E,Aebersold R. A statistical model
for identifying proteins by tandem mass spectrometry. Anal Chem
75: 46464658, 2003. doi:10.1021/ac0341261.
60. Kr
uger J,Rehmsmeier M. RNAhybrid: microRNA target prediction
easy, fast and exible. Nucleic Acids Res 34: W451W454, 2006.
doi:10.1093/nar/gkl243.
61. Rehmsmeier M,Steffen P,Hochsmann M,Giegerich R. Fast and
effective prediction of microRNA/target duplexes. RNA 10: 1507
1517, 2004. doi:10.1261/rna.5248604.
62. Vechetti IJ Jr,Peck BD,Wen Y,Walton RG,Valentino TR,Alimov
AP,Dungan CM,Van Pelt DW,von Walden F,Alkner B,Peterson
CA,McCarthy JJ. Mechanical overload-induced muscle-derived
extracellular vesicles promote adipose tissue lipolysis. FASEB J 35:
e21644, 2021. doi:10.1096/fj.202100242R.
63. Roy RR,Edgerton V. Response of mouse plantaris muscle to func-
tional overload: comparison with rat and cat. Comp Biochem Physiol
APhysiol111: 569575, 1995. doi:10.1016/0300-9629(95)00062-c.
64. Murach KA,McCarthy JJ,Peterson CA,Dungan CM. Making mice
mighty: recent advances in translational models of load-induced
muscle hypertrophy. J Appl Physiol (1985) 129: 516521, 2020.
doi:10.1152/japplphysiol.00319.2020.
65. Ruest LB,Marcotte R,Wang E. Peptide elongation factor eEF1A-2/
S1 expression in cultured differentiated myotubes and its protective
effect against caspase-3-mediated apoptosis. JBiolChem277:
54185425, 2002. doi:10.1074/jbc.M110685200.
66. Piazzi M,Bavelloni A,Faenza I,Blalock W,Urbani A,DAguanno S,
Fiume R,Ramazzotti G,Maraldi NM,Cocco L. eEF1A phosphoryla-
tion in the nucleus of insulin-stimulated C2C12 myoblasts. Mol Cell
Proteomics 9: 27192728, 2010. doi:10.1074/mcp.M110.003152.
67. Vislovukh AA,Groisman IS,Elskaya AV,Negrutskii BS,Polesskaya
AN. Transcriptional and post-transcriptional control of eEF1A2 expres-
sion during myoblast differentiation. Biopolymers Cell 28: 456460,
2012. doi:10.7124/bc.000136.
68. Baker N,Wade S,Triolo M,Girgis J,Chwastek D,Larrigan S,Feige
P,Fujita R,Crist C,Rudnicki MA,Burelle Y,Khacho M. The mito-
chondrial protein OPA1 regulates the quiescent state of adult muscle
stem cells. Cell Stem Cell 29: 13151332.e9, 2022. doi:10.1016/j.
stem.2022.07.010.
69. Sin J,Andres AM,Taylor DJ,Weston T,Hiraumi Y,Stotland A,Kim
BJ,Huang C,Doran KS,Gottlieb RA. Mitophagy is required for
miR-16 IN MYOGENIC CELLS
C1108 AJP-Cell Physiol doi:10.1152/ajpcell.00071.2023 www.ajpcell.org
Downloaded from journals.physiology.org/journal/ajpcell at Univ of Arkansas Fayetteville (130.184.252.076) on May 18, 2023.
mitochondrial biogenesis and myogenic differentiation of C2C12 myo-
blasts. Autophagy 12: 369380, 2016. doi:10.1080/15548627.2015.
1115172.
70. Yue F,Bi P,Wang C,Shan T,Nie Y,Ratliff TL,Gavin TP,Kuang S.
Pten is necessary for the quiescence and maintenance of adult muscle
stem cells. Nat Commun 8: 14328, 2017. doi:10.1038/ncomms14328.
71. Yue F,Bi P,Wang C,Li J,Liu X,Kuang S. Conditional loss of Pten in
myogenic progenitors leads to postnatal skeletal muscle hypertro-
phy but age-dependent exhaustion of satellite cells. Cell Rep 17:
23402353, 2016. doi:10.1016/j.celrep.2016.11.002.
72. Langdon CG,Gadek KE,Garcia MR,Evans MK,Reed KB,Bush M,
Hanna JA,Drummond CJ,Maguire MC,Leavey PJ,Finkelstein D,
Jin H,Schreiner PA,Rehg JE,Hatley ME. Synthetic essentiality
between PTEN and core dependency factor PAX7 dictates rhabdo-
myosarcoma identity. Nat Commun 12: 5520, 2021. doi:10.1038/
s41467-021-25829-4.
73. Filipowicz W,Bhattacharyya SN,Sonenberg N. Mechanisms of
post-transcriptional regulation by microRNAs: are the answers in
sight? Nat Rev Genet 9: 102114, 2008. doi:10.1038/nrg2290.
74. Rando TA,Blau HM. Primary mouse myoblast purication, charac-
terization, and transplantation for cell-mediated gene therapy. JCell
Biol 125: 12751287, 1994. doi:10.1083/jcb.125.6.1275.
75. von Walden F,Rea M,Mobley CB,Fondufe-Mittendorf Y,McCarthy
JJ,Peterson CA,Murach KA. The myonuclear DNA methylome in
response to an acute hypertrophic stimulus. Epigenetics 15: 11511162,
2020. doi:10.1080/15592294.2020.1755581.
76. Murach KA,White SH,Wen Y,Ho A,Dupont-Versteegden EE,
McCarthy JJ,Peterson CA. Differential requirement for satellite cells
during overload-induced muscle hypertrophy in growing versus
mature mice. Skelet Muscle 7: 14, 2017. doi:10.1186/s13395-017-0132-z.
77. Stanseld BN,Brown AD,Stewart CE,Burniston JG. Dynamic prol-
ing of protein mole synthesis rates during C2C12 myoblast differentia-
tion. Proteomics 21: 2000071, 2021. doi:10.1002/pmic.202000071.
78. Brown AD,Stewart CE,Burniston JG. Degradation of ribosomal
and chaperone proteins is attenuated during the differentiation of
replicatively aged C2C12 myoblasts. JCachexiaSarcopeniaMuscle
13: 25622575, 2022. doi:10.1002/jcsm.13034.
79. Murach KA,Dungan CM,von Walden F,Wen Y. Epigenetic evi-
dence for distinct contributions of resident and acquired myonuclei
during long-term exercise adaptation using timed in vivo myonuclear
labeling. Am J Physiol Cell Physiol 322: C86C93, 2022. doi:10.1152/
ajpcell.00358.2021.
80. Biasini A,Abdulkarim B,de Pretis S,Tan JY,Arora R,Wischnewski
H,Dreos R,Pelizzola M,Ciaudo C,Marques AC. Translation is
required for miRNA-dependent decay of endogenous transcripts.
EMBO J 40: e104569, 2021. doi:10.15252/embj.2020104569.
81. Xu X,Ji S,Li W,Yi B,Li H,Zhang H,Ma W. LncRNA H19 promotes
the differentiation of bovine skeletal muscle satellite cells by sup-
pressing Sirt1/FoxO1. Cell Mol Biol Lett 22: 10, 2017. doi:10.1186/
s11658-017-0040-6.
82. Li J,Su T,Zou C,Luo W,Shi G,Chen L,Fang C,Li C. Long non-cod-
ing RNA H19 regulates porcine satellite cell differentiation through
miR-140-5p/SOX4 and DBN1. Front Cell Dev Biol 8: 518724, 2020.
doi:10.3389/fcell.2020.518724.
83. Li J,Zhao W,Li Q,Huang Z,Shi G,Li C. Long non-coding RNA H19
Promotes porcine satellite cell differentiation by interacting with
TDP43. Genes 11: 259, 2020. doi:10.3390/genes11030259.
84. Dey BK,Pfeifer K,Dutta A. The H19 long noncoding RNA gives rise
to microRNAs miR-675-3p and miR-675-5p to promote skeletal mus-
cle differentiation and regeneration. Genes Dev 28: 491501, 2014.
doi:10.1101/gad.234419.113.
85. Chen X,He L,Zhao Y,Li Y,Zhang S,Sun K,So K,Chen F,Zhou L,
Lu L,Wang L,Zhu X,Bao X,Esteban MA,Nakagawa S,Prasanth
KV,Wu Z,Sun H,Wang H. Malat1 regulates myogenic differentiation
and muscle regeneration through modulating MyoD transcriptional
activity. Cell Discov 3: 17002, 2017. doi:10.1038/celldisc.2017.2.
86. Millay DP,ORourke JR,Sutherland LB,Bezprozvannaya S,
Shelton JM,Bassel-Duby R,Olson EN. Myomaker is a membrane
activator of myoblast fusion and muscle formation. Nature 499: 301
305, 2013. doi:10.1038/nature12343.
87. McKellar DW,Walter LD,Song LT,Mantri M,Wang MF,De
Vlaminck I,Cosgrove BD. Large-scale integration of single-cell tran-
scriptomic data captures transitional progenitor states in mouse skele-
tal muscle regeneration. Commun Biol 4: 1280, 2021. doi:10.1038/
s42003-021-02810-x.
88. Xue Q,Zhang G,Li T,Ling J,Zhang X,Wang J. Transcriptomic pro-
le of leg muscle during early growth in chicken. PLoS One 12:
e0173824, 2017. doi:10.1371/journal.pone.0173824.
89. Maire P,Dos Santos M,Madani R,Sakakibara I,Viaut C,Wurmser
M. Myogenesis control by SIX transcriptional complexes. Semin Cell
Dev Biol 104: 5164, 2020. doi:10.1016/j.semcdb.2020.03.003.
90. Randrianarison-Huetz V,Papaefthymiou A,Herledan G,Noviello
C,Faradova U,Collard L,Pincini A,Schol E,Decaux JF,Maire P,
Vassilopoulos S,Sotiropoulos A. Srf controls satellite cell fusion
through the maintenance of actin architecture. JCellBiol217: 685
700, 2018. doi:10.1083/jcb.201705130.
91. Al Tanoury Z,Rao J,Tassy O,Gobert B,Gapon S,Garnier JM,
Wagner E,Hick A,Hall A,Gussoni E,Pourqui
eO.Differentiation of
the human PAX7-positive myogenic precursors/satellite cell lineage in
vitro. Development 147: dev187344, 2020. doi:10.1242/dev.187344.
92. Wang J,Broer T,Chavez T,Zhou CJ,Tran S,Xiang Y,Khodabukus
A,Diao Y,Bursac N. Myoblast deactivation within engineered
human skeletal muscle creates a transcriptionally heterogeneous
population of quiescent satellite-like cells. Biomaterials 284: 121508,
2022. doi:10.1016/j.biomaterials.2022.121508.
93. Qiu X,Hill A,Packer J,Lin D,Ma YA,Trapnell C. Single-cell mRNA
quantication and differential analysis with Census. Nat Methods 14:
309315, 2017. doi:10.1038/nmeth.4150.
94. Hori S,Hiramuki Y,Nishimura D,Sato F,Sehara-Fujisawa A. PDH-
mediated metabolic ow is critical for skeletal muscle stem cell dif-
ferentiation and myotube formation during regeneration in mice.
FASEB J 33: 80948109, 2019. doi:10.1096/fj.201802479R.
95. Kahns S,Knudsen C,Clark B,Lund A,Kristensen P,Cavallius J,
Merrick W. The elongation factor 1 A-2 isoform from rabbit: cloning
of the cDNA and characterization of the protein. Nucleic Acids Res
26: 18841890, 1998. doi:10.1093/nar/26.8.1884.
96. Hudson NJ,Lyons RE,Reverter A,Greenwood PL,Dalrymple BP.
Inferring the in vivo cellular program of developing bovine skeletal
muscle from expression data. Gene Expr Patterns 13: 109125, 2013.
doi:10.1016/j.gep.2013.02.001.
97. P
eladeau C,Adam N,Bronicki LM,Coriati A,Thabet M,Al-
Rewashdy H,Vanstone J,Mears A,Renaud JM,Holcik M,Jasmin
BJ. Identication of therapeutics that target eEF1A2 and upregulate
utrophin A translation in dystrophic muscles. Nat Commun 11: 1990,
2020. doi:10.1038/s41467-020-15971-w.
98. Miura P,Coriati A,B
elanger G,De Repentigny Y,Lee J,Kothary R,
Holcik M,Jasmin B. The utrophin A 5 0-UTR drives cap-independent
translation exclusively in skeletal muscles of transgenic mice and inter-
acts with eEF1A2. Hum Mol Genet 19: 12111220, 2010. doi:10.1093/
hmg/ddp591.
99. Lyu Y,Jia W,Wu Y,Zhao X,Xia Y,Guo X,Kang J. Cpmer: a new
conserved eEF1A2-binding partner that regulates Eomes translation
and cardiomyocyte differentiation. Stem Cell Reports 17: 11541169,
2022. doi:10.1016/j.stemcr.2022.03.006.
100. Duan S,Yuan G,Liu X,Ren R,Li J,Zhang W,Wu J,Xu X,Fu L,Li Y,
Yang J,Zhang W,Bai R,Yi F,Suzuki K,Gao H,Esteban CR,Zhang C,
Izpisua Belmonte JC,Chen Z,Wang X,Jiang T,Qu J,Tang F,Liu G-
H. PTEN deciency reprogrammes human neural stem cells towards a
glioblastoma stem cell-like phenotype. Nat Commun 6: 10068, 2015.
doi:10.1038/ncomms10068.
miR-16 IN MYOGENIC CELLS
AJP-Cell Physiol doi:10.1152/ajpcell.00071.2023 www.ajpcell.org C1109
Downloaded from journals.physiology.org/journal/ajpcell at Univ of Arkansas Fayetteville (130.184.252.076) on May 18, 2023.
... Manipulating miR-16 on myofiber-associated satellite cells in vitro via hairpin inhibitors affects their fate [171]. Thus, we recently focused on the role of miR-16 in myogenic cells [172]. Proteomic profiling of C2C12 myoblasts after miR-16 knockdown suggested that miR-16 may influence myogenic cell differentiation through upregulation of EEF1A2, OPA1, and PTEN [172]; all of these factors have defined roles in controlling satellite cell fate [173][174][175][176][177][178][179][180]. ...
... Thus, we recently focused on the role of miR-16 in myogenic cells [172]. Proteomic profiling of C2C12 myoblasts after miR-16 knockdown suggested that miR-16 may influence myogenic cell differentiation through upregulation of EEF1A2, OPA1, and PTEN [172]; all of these factors have defined roles in controlling satellite cell fate [173][174][175][176][177][178][179][180]. Recent advances in single cell and spatial sequencing technology will enable the characterization of small RNAs in individual satellite cells at different J o u r n a l P r e -p r o o f 14 phases during adult muscle adaptation [181]. ...
Article
Satellite cells are bona fide muscle stem cells that are indispensable for successful post-natal muscle growth and regeneration after severe injury. These cells also participate in adult muscle adaptation in several capacities. microRNA (miRNA) are post-transcriptional regulators of mRNA that are implicated in several aspects of stem cell function. There is evidence to suggest that miRNAs affect satellite cell behavior in vivo and myogenic progenitor behavior in vitro, but the role of miRNAs in adult skeletal muscle satellite cells is less studied. In this review, we provide evidence for how miRNAs control satellite cell behavior with emphasis on satellite cells of adult muscle in vivo. We first outline how miRNAs are indispensable for satellite cell viability and control the phases of myogenesis. Next, we discuss the interplay between miRNAs and myogenic cell redox status, senescence, and communication to other muscle-resident cells during muscle adaptation. Results from recent satellite cell miRNA profiling studies are also summarized. In vitro experiments in primary myogenic cells and cell lines have been invaluable for exploring the influence of miRNAs, but we identify a need for novel genetic tools to further interrogate how miRNAs control satellite cell behavior in adult skeletal muscle in vivo.
... MicroRNAs (miRNAs) are $20 nucleotides in length, are noncoding, and act to inhibit the translation of mRNAs in a sequence-specific fashion via the miRNA-induced silencing complex (miRISC) (683,684 [693][694][695][696], and recent in vitro and rodent evidence suggests that miR-16 is lower in mechanically overloaded muscle and that this may lead to the derepression of genes involved in myoblast differentiation and ribosome biogenesis (697). Although these data have been informative, a 2019 study by Vechetti et al. (698) tempers enthusiasm in this area. ...
Article
Full-text available
Mechanisms underlying mechanical overload-induced skeletal muscle hypertrophy have been extensively researched since the landmark report by Morpurgo (1897) of "work-induced hypertrophy" in dogs that were treadmill-trained. Much of the pre-clinical rodent and human resistance training research to date supports that involved mechanisms include enhanced mammalian/mechanistic target of rapamycin complex 1 (mTORC1) signaling, an expansion in translational capacity through ribosome biogenesis, increased satellite cell abundance and myonuclear accretion, and post-exercise elevations in muscle protein synthesis rates. However, several lines of past and emerging evidence suggest additional mechanisms that feed into or are independent of these processes are also involved. This review will first provide a historical account as to how mechanistic research into skeletal muscle hypertrophy has progressed. A comprehensive list of mechanisms associated with skeletal muscle hypertrophy is then outlined and areas of disagreement involving these mechanisms are presented. Finally, future research directions involving many of the discussed mechanisms will be proposed.
... Additionally, miR-100-5p has a highly conservative sequence among different species (mouse, human, chicken, pig, and cattle). These results suggest that miR-100-5p might be a muscle-associated miRNA, similar to previously reported muscle-related miRNAs [39][40][41]. ...
Article
Full-text available
MicroRNAs (miRNAs) are endogenous small non-coding RNAs that play crucial regulatory roles in many biological processes, including the growth and development of skeletal muscle. miRNA-100-5p is often associated with tumor cell proliferation and migration. This study aimed to uncover the regulatory mechanism of miRNA-100-5p in myogenesis. In our study, we found that the miRNA-100-5p expression level was significantly higher in muscle tissue than in other tissues in pigs. Functionally, this study shows that miR-100-5p overexpression significantly promotes the proliferation and inhibits the differentiation of C2C12 myoblasts, whereas miR-100-5p inhibition results in the opposite effects. Bioinformatic analysis predicted that Trib2 has potential binding sites for miR-100-5p at the 3′UTR region. A dual-luciferase assay, qRT-qPCR, and Western blot confirmed that Trib2 is a target gene of miR-100-5p. We further explored the function of Trib2 in myogenesis and found that Trib2 knockdown markedly facilitated proliferation but suppressed the differentiation of C2C12 myoblasts, which is contrary to the effects of miR-100-5p. In addition, co-transfection experiments demonstrated that Trib2 knockdown could attenuate the effects of miR-100-5p inhibition on C2C12 myoblasts differentiation. In terms of the molecular mechanism, miR-100-5p suppressed C2C12 myoblasts differentiation by inactivating the mTOR/S6K signaling pathway. Taken together, our study results indicate that miR-100-5p regulates skeletal muscle myogenesis through the Trib2/mTOR/S6K signaling pathway.
Article
Full-text available
In women who are getting older, the quantity and quality of their follicles or oocytes and decline. This is characterized by decreased ovarian reserve function (DOR), fewer remaining oocytes, and lower quality oocytes. As more women choose to delay childbirth, the decline in fertility associated with age has become a significant concern for modern women. The decline in oocyte quality is a key indicator of ovarian aging. Many studies suggest that age-related changes in oocyte energy metabolism may impact oocyte quality. Changes in oocyte energy metabolism affect adenosine 5'-triphosphate (ATP) production, but how related products and proteins influence oocyte quality remains largely unknown. This review focuses on oocyte metabolism in age-related ovarian aging and its potential impact on oocyte quality, as well as therapeutic strategies that may partially influence oocyte metabolism. This research aims to enhance our understanding of age-related changes in oocyte energy metabolism, and the identification of biomarkers and treatment methods.
Article
Disuse-induced muscle atrophy is a common clinical problem observed mainly in older adults, intensive care units patients, or astronauts. Previous studies presented biological sex divergence in progression of disuse-induced atrophy along with differential changes in molecular mechanisms possibly underlying muscle atrophy. The aim of this study was to perform transcriptomic profiling of male and female mice during the onset and progression of unloading disuse-induced atrophy. Male and female mice underwent hindlimb unloading (HU) for 24, 48, 72 and 168h (n=8/group). Muscles were weighed for each cohort and gastrocnemius was used for RNA-sequencing analysis. Females exhibited muscle loss as early as 24h of HU, while males after 168h of HU. In males, pathways related to proteasome degradation were upregulated throughout 168-h HU, while in females these pathways were upregulated up to 72-h HU. Lcn2, a gene contributing to regulation of myogenesis, was upregulated by 6.46-19.86-fold across all time points in females only. A reverse expression of Fosb, a gene related to muscle degeneration, was observed between males (4.27-fold up) and females (4.57-fold down) at 24-h HU. Mitochondrial pathways related to TCA cycle were highly downregulated at 168h of HU in males, while in females this downregulation was less pronounced. Collagen-related pathways were consistently downregulated throughout 168-h HU only in females, suggesting a potential biological sex-specific protective mechanism against disuse-induced fibrosis. In conclusion, females may have protection against HU-induced skeletal muscle mitochondrial degeneration and fibrosis through transcriptional mechanisms, although they may be more vulnerable to HU-induced muscle wasting compared to males.
Article
Full-text available
Most cells in the body are mononuclear whereas skeletal muscle fibres are uniquely multinuclear. The nuclei of muscle fibres (myonuclei) are usually situated peripherally which complicates the equitable distribution of gene products. Myonuclear abundance can also change under conditions such as hypertrophy and atrophy. Specialised zones in muscle fibres have different functions and thus distinct synthetic demands from myonuclei. The complex structure and regulatory requirements of multinuclear muscle cells understandably led to the hypothesis that myonuclei govern defined ‘domains’ to maintain homeostasis and facilitate adaptation. The purpose of this review is to provide historical context for the myonuclear domain and evaluate its veracity with respect to mRNA and protein distribution resulting from myonuclear transcription. We synthesise insights from past and current in vitro and in vivo genetically modified models for studying the myonuclear domain under dynamic conditions. We also cover the most contemporary knowledge on mRNA and protein transport in muscle cells. Insights from emerging technologies such as single myonuclear RNA‐sequencing further inform our discussion of the myonuclear domain. We broadly conclude: (1) the myonuclear domain can be flexible during muscle fibre growth and atrophy, (2) the mechanisms and role of myonuclear loss and motility deserve further consideration, (3) mRNA in muscle is actively transported via microtubules and locally restricted, but proteins may travel far from a myonucleus of origin and (4) myonuclear transcriptional specialisation extends beyond the classic neuromuscular and myotendinous populations. A deeper understanding of the myonuclear domain in muscle may promote effective therapies for ageing and disease. image
Article
Full-text available
Myc is a powerful transcription factor implicated in epigenetic reprogramming, cellular plasticity, and rapid growth as well as tumorigenesis. Cancer in skeletal muscle is extremely rare despite marked and sustained Myc induction during loading-induced hypertrophy. Here, we investigated global, actively transcribed, stable, and myonucleus-specific transcriptomes following an acute hypertrophic stimulus in mouse plantaris. With these datasets we define global and Myc-specific dynamics at the onset of mechanical overload-induced muscle fiber growth. Data collation across analyses reveals an under-appreciated role for the muscle fiber in extracellular matrix remodeling during adaptation, along with the contribution of mRNA stability to epigenetic-related transcript levels in muscle. We also identify Runx1 and Ankrd1 (Marp1) as abundant myonucleus-enriched loading-induced genes. We observed that a strong induction of cell cycle regulators including Myc occurs with mechanical overload in myonuclei. Additionally, in vivo Myc-controlled gene expression in the plantaris was defined using a genetic muscle fiber-specific doxycycline-inducible Myc-overexpression model. We determined Myc is implicated in numerous aspects of gene expression during early-phase muscle fiber growth. Specifically, brief induction of Myc protein in muscle represses Reverbα, Reverbβ, and Myh2 while increasing Rpl3, recapitulating gene expression in myonuclei during acute overload. Experimental, comparative, and in silico analyses place Myc at the center of a stable and actively transcribed, loading-responsive, muscle fiber-localized regulatory hub. Collectively, our experiments are a roadmap for understanding global and Myc-mediated transcriptional networks that regulate rapid remodeling in post-mitotic cells. We provide open webtools for exploring the five RNA-sequencing datasets as a resource to the field.
Article
Full-text available
Quiescence regulation is essential for adult stem cell maintenance and sustained regeneration. Our studies uncovered that physiological changes in mitochondrial shape regulate the quiescent state of adult muscle stem cells (MuSCs). We show that MuSC mitochondria rapidly fragment upon an activation stimulus, via systemic HGF/mTOR, to drive the exit from deep quiescence. Deletion of the mitochondrial fusion protein OPA1 and mitochondrial fragmentation transitions MuSCs into G-alert quiescence, causing premature activation and depletion upon a stimulus. OPA1 loss activates a glutathione (GSH)-redox signaling pathway promoting cell-cycle progression, myogenic gene expression, and commitment. MuSCs with chronic OPA1 loss, leading to mitochondrial dysfunction, continue to reside in G-alert but acquire severe cell-cycle defects. Additionally, we provide evidence that OPA1 decline and impaired mitochondrial dynamics contribute to age-related MuSC dysfunction. These findings reveal a fundamental role for OPA1 and mitochondrial dynamics in establishing the quiescent state and activation potential of adult stem cells.
Article
Full-text available
Background: Cell assays are important for investigating the mechanisms of ageing, including losses in protein homeostasis and 'proteostasis collapse'. We used novel isotopic labelling and proteomic methods to investigate protein turnover in replicatively aged (>140 population doublings) murine C2C12 myoblasts that exhibit impaired differentiation and serve as a model for age-related declines in muscle homeostasis. Methods: The Absolute Dynamic Profiling Technique for Proteomics (Proteo-ADPT) was used to investigate proteostasis in young (passage 6-10) and replicatively aged (passage 48-50) C2C12 myoblast cultures supplemented with deuterium oxide (D2 O) during early (0-24 h) or late (72-96 h) periods of differentiation. Peptide mass spectrometry was used to quantify the absolute rates of abundance change, synthesis and degradation of individual proteins. Results: Young cells exhibited a consistent ~25% rise in protein accretion over the 96-h experimental period. In aged cells, protein accretion increased by 32% (P < 0.05) during early differentiation, but then fell back to baseline levels by 96-h. Proteo-ADPT encompassed 116 proteins and 74 proteins exhibited significantly (P < 0.05, FDR < 5% interaction between age × differentiation stage) different changes in abundance between young and aged cells at early and later periods of differentiation, including proteins associated with translation, glycolysis, cell-cell adhesion, ribosomal biogenesis, and the regulation of cell shape. During early differentiation, heat shock and ribosomal protein abundances increased in aged cells due to suppressed degradation rather than heightened synthesis. For instance, HS90A increased at a rate of 10.62 ± 1.60 ng/well/h in aged which was significantly greater than the rate of accretion (1.86 ± 0.49 ng/well/h) in young cells. HS90A synthesis was similar in young (21.23 ± 3.40 ng/well/h) and aged (23.69 ± 1.13 ng/well/h), but HS90A degradation was significantly (P = 0.05) greater in young (19.37 ± 2.93 ng/well/h) versus aged (13.06 ± 0.76 ng/well/h) cells. During later differentiation the HS90A degradation (8.94 ± 0.38 ng/well/h) and synthesis (7.89 ± 1.28 ng/well/h) declined and were significantly less than the positive net balance between synthesis and degradation (synthesis = 28.14 ± 3.70 ng/well/h vs. degradation = 21.49 ± 3.13 ng/well/h) in young cells. Conclusions: Our results suggest a loss of proteome quality as a precursor to the lack of fusion of aged myoblasts. The quality of key chaperone proteins, including HS90A, HS90B and HSP7C was reduced in aged cells and may account for the disruption to cell signalling required for the later stages of differentiation and fusion.
Article
Full-text available
Previous studies have shown that eukaryotic elongation factor 1A2 (eEF1A2) serves as an essential heart-specific translation elongation element and that its mutation or knockout delays heart development and causes congenital heart disease and death among species. However, the function and regulatory mechanisms of eEF1A2 in mammalian heart development remain largely unknown. Here we identified the long noncoding RNA (lncRNA) Cpmer (cytoplasmic mesoderm regulator), which interacted with eEF1A2 to co-regulate differentiation of mouse and human embryonic stem cell-derived cardiomyocytes. Mechanistically, Cpmer specifically recognized Eomes mRNA by RNA-RNA pairing and facilitated binding of eEF1A2 with Eomes mRNA, guaranteeing Eomes mRNA translation and cardiomyocyte differentiation. Our data reveal a novel functionally conserved lncRNA that can specifically regulate Eomes translation and cardiomyocyte differentiation, which broadens our understanding of the mechanism of lncRNA involvement in the subtle translational regulation of eEF1A2 during mammalian heart development.
Article
Full-text available
Molecular interactions are key drivers of biological function. Providing interaction resources to the research community is important since they allow functional interpretation and network-based analysis of molecular data. ConsensusPathDB (http://consensuspathdb.org) is a meta-database combining interactions of diverse types from 31 public resources for humans, 16 for mice and 14 for yeasts. Using ConsensusPathDB, researchers commonly evaluate lists of genes, proteins and metabolites against sets of molecular interactions defined by pathways, Gene Ontology and network neighborhoods and retrieve complex molecular neighborhoods formed by heterogeneous interaction types. Furthermore, the integrated protein–protein interaction network is used as a basis for propagation methods. Here, we present the 2022 update of ConsensusPathDB, highlighting content growth, additional functionality and improved database stability. For example, the number of human molecular interactions increased to 859 848 connecting 200 499 unique physical entities such as genes/proteins, metabolites and drugs. Furthermore, we integrated regulatory datasets in the form of transcription factor–, microRNA– and enhancer–gene target interactions, thus providing novel functionality in the context of overrepresentation and enrichment analyses. We specifically emphasize the use of the integrated protein–protein interaction network as a scaffold for network inferences, present topological characteristics of the network and discuss strengths and shortcomings of such approaches.
Article
Full-text available
Muscle fibers are syncytial post-mitotic cells that can acquire exogenous nuclei from resident muscle stem cells, called satellite cells. Myonuclei are added to muscle fibers by satellite cells during conditions such as load-induced hypertrophy. It is difficult to dissect the molecular contributions of resident versus satellite cell-derived myonuclei during adaptation due to the complexity of labeling distinct nuclear populations in multinuclear cells without label transference between nuclei. To sidestep this barrier, we utilized a genetic mouse model where myonuclear DNA can be specifically and stably labeled via non-constitutive H2B-GFP at any point in the lifespan. Resident myonuclei (Mn) were GFP-tagged in vivo before eight weeks of progressive weighted wheel running (PoWeR) in adult mice (>4-month-old). Resident+satellite cell-derived myonuclei (Mn+SC Mn) were labeled at the end of PoWeR in a separate cohort. Following myonuclear isolation, promoter DNA methylation profiles acquired with low-input RRBS were compared to deduce epigenetic contributions of satellite cell-derived myonuclei during adaptation. Resident myonuclear DNA has hypomethylated promoters in genes related to protein turnover, while the addition of satellite cell-derived myonuclei shifts myonuclear methylation profiles to favor transcription factor regulation and cell-cell signaling. By comparing myonucleus-specific methylation profiling to previously published single-nucleus transcriptional analysis in the absence (Mn) versus presence of satellite cells (Mn+SC Mn) with PoWeR, we provide evidence that satellite cell-derived myonuclei may preferentially supply ribosomal proteins to growing myofibers and retain an epigenetic "memory" of prior stem cell identity. These data offer insights on distinct epigenetic myonuclear characteristics and contributions during adult muscle growth.
Article
Full-text available
Skeletal muscle repair is driven by the coordinated self-renewal and fusion of myogenic stem and progenitor cells. Single-cell gene expression analyses of myogenesis have been hampered by the poor sampling of rare and transient cell states that are critical for muscle repair, and do not inform the spatial context that is important for myogenic differentiation. Here, we demonstrate how large-scale integration of single-cell and spatial transcriptomic data can overcome these limitations. We created a single-cell transcriptomic dataset of mouse skeletal muscle by integration, consensus annotation, and analysis of 23 newly collected scRNAseq datasets and 88 publicly available single-cell (scRNAseq) and single-nucleus (snRNAseq) RNA-sequencing datasets. The resulting dataset includes more than 365,000 cells and spans a wide range of ages, injury, and repair conditions. Together, these data enabled identification of the predominant cell types in skeletal muscle, and resolved cell subtypes, including endothelial subtypes distinguished by vessel-type of origin, fibro-adipogenic progenitors defined by functional roles, and many distinct immune populations. The representation of different experimental conditions and the depth of transcriptome coverage enabled robust profiling of sparsely expressed genes. We built a densely sampled transcriptomic model of myogenesis, from stem cell quiescence to myofiber maturation, and identified rare, transitional states of progenitor commitment and fusion that are poorly represented in individual datasets. We performed spatial RNA sequencing of mouse muscle at three time points after injury and used the integrated dataset as a reference to achieve a high-resolution, local deconvolution of cell subtypes. We also used the integrated dataset to explore ligand-receptor co-expression patterns and identify dynamic cell-cell interactions in muscle injury response. We provide a public web tool to enable interactive exploration and visualization of the data. Our work supports the utility of large-scale integration of single-cell transcriptomic data as a tool for biological discovery. David McKellar et al. integrate single-cell and -nuclear transcriptomic analyses of mouse skeletal muscle in homeostatic conditions or following injury. The resulting transcriptomic model of myogenesis identified rare, transitional states and cell subtypes that are poorly represented in individual datasets, providing a valuable resource for the field.
Article
Full-text available
Cancer cachexia (CC) results in impaired muscle function and quality of life and is the primary cause of death for ~20-30% of cancer patients. We demonstrated mitochondrial degeneration as a precursor to CC in male mice, however, if such alterations occur in females is currently unknown. The purpose of this study was to elucidate muscle alterations in CC development in female tumor-bearing mice. 60 female C57BL/6J mice were injected with PBS or Lewis Lung Carcinoma at 8-week age, and tumors developed for 1, 2, 3, or 4 weeks to assess the time course of cachectic development. In vivo muscle contractile function, protein fractional synthetic rate (FSR), protein turnover, and mitochondrial health were assessed. 3- and 4-week tumor-bearing mice displayed a dichotomy in tumor growth and were reassigned to High Tumor (HT) and Low Tumor (LT) groups. HT mice exhibited lower soleus, TA, and fat weights compared to PBS. HT mice showed lower peak isometric torque and slower one-half relaxation time compared to PBS. HT mice had lower FSR compared to PBS while E3 ubiquitin ligases were greater in HT compared to other groups. Bnip3 (mitophagy) and pMitoTimer red puncta (mitochondrial degeneration) were greater in HT while Pgc1α1 and Tfam (mitochondrial biogenesis) were lower in HT compared to PBS. We demonstrate alterations in female tumor-bearing mice where HT exhibited greater protein degradation, impaired muscle contractility, and mitochondrial degeneration compared to other groups. Our data provide novel evidence for a distinct cachectic development in tumor-bearing female mice compared to previous male studies.
Article
Satellite cells (SCs), the adult Pax7-expressing stem cells of skeletal muscle, are essential for muscle repair. However, in vitro investigations of SC function are challenging due to isolation-induced SC activation, loss of native quiescent state, and differentiation to myoblasts. In the present study, we optimized methods to deactivate in vitro expanded human myoblasts within a 3D culture environment of engineered human skeletal muscle tissues (“myobundles”). Immunostaining and gene expression analyses revealed that a fraction of myoblasts within myobundles adopted a quiescent phenotype (3D-SCs) characterized by increased Pax7 expression, cell cycle exit, and activation of Notch signaling. Similar to native SCs, 3D-SC quiescence is regulated by Notch and Wnt signaling while loss of quiescence and reactivation of 3D-SCs can be induced by growth factors including bFGF. Myobundle injury with a bee toxin, melittin, induces robust myofiber fragmentation, functional decline, and 3D-SC proliferation. By applying single cell RNA-sequencing (scRNA-seq), we discover the existence of two 3D-SC subpopulations (quiescent and activated), identify deactivation-associated gene signature using trajectory inference between 2D myoblasts and 3D-SCs, and characterize the transcriptomic changes within reactivated 3D-SCs in response to melittin-induced injury. These results demonstrate the ability of an in vitro engineered 3D human skeletal muscle environment to support the formation of a quiescent and heterogeneous SC population recapitulating several aspects of the native SC phenotype, and provide a platform for future studies of human muscle regeneration and disease-associated SC dysfunction.