ArticlePDF Available

Myostatin modulates adipogenesis to generate adipocytes with favorable metabolic effects

Authors:

Abstract and Figures

A pluripotent cell line, C3H10T1/2, is induced to undergo adipogenesis by a mixture of factors that includes a glucocorticoid such as dexamethasone. We found that expression of myostatin (MSTN), a TGF-beta family member extensively studied in muscle, was induced by dexamethasone under those differentiation conditions. Moreover, MSTN could substitute for dexamethasone in the adipogenesis mixture. However, the adipocytes induced by MSTN in both cell culture and transgenic mice were small and expressed markers characteristic of immature adipocytes. These adipocytes exhibited cell-autonomous increases in insulin sensitivity and glucose oxidation. In mice, these effects produced elevated systemic insulin sensitivity and resistance to diet-induced obesity. Modulation of the final stages of adipogenesis may provide a novel approach to understanding and treating metabolic disease.
aP2-MSTN mice have a favorable metabolic profile and are resistant to obesity. ( A ) The transgene (Tg) construct was created by subcloning the full-length MSTN cDNA behind an aP2 promoter. Transgenic mice had a 10-fold overexpression of MSTN as detected by quantitative real-time PCR amplification of cDNA generated from the adipose tissue of WT and transgenic mice. Primers P1 and P2 were used to amplify a transgene-specific cDNA, whereas primers P1 and P3 were used to compare total MSTN expression levels. Quantitative PCR amplification of hypoxanthine-guanine phosphoribosyl transferase (HPRT) was used to control for cDNA concentration between samples. ( B ) Glucose tolerance tests were performed on both lean mice ( Upper ) and mice on a high-fat diet ( Lower ). Blood sugar levels were checked after an overnight fast (time ϭ 0) and then 2 g ͞ kg of glucose was injected i.p. After the glucose bolus, tail blood sugar levels were checked at each time point. At the 60-min time point, all WT animals on the high-fat diet had blood sugars above the upper limit of the meter, therefore the upper limit value (600 mg ͞ dl) was used at that time point. aP2-MSTN mice were more insulin sensitive than their WT littermates on both regular ( P ϭ 0.0061) and high-fat chow ( P ϭ 0.0025). ( C ) aP2-MSTN mice and their WT littermates were placed on a high-fat diet for 7 weeks. Transgenic mice ( Lower Right ) were resistant to the obesity that developed in WT animals ( Lower Left ). ( D ) aP2-MSTN mice had lower fasting glucose, insulin, and triglyceride levels than their WT littermates. Error bars represent SDs.
… 
Content may be subject to copyright.
Myostatin modulates adipogenesis to generate
adipocytes with favorable metabolic effects
Brian J. Feldman*
, Ryan S. Streeper
, Robert V. Farese, Jr.
द
, and Keith R. Yamamoto
Departments of *Pediatrics,
Cellular and Molecular Pharmacology,
§
Medicine, and
Biochemistry and Biophysics, University of California,
San Francisco, CA 94143; and
Gladstone Institute of Cardiovascular Disease, San Francisco, CA 94143
Contributed by Keith R. Yamamoto, August 28, 2006
A pluripotent cell line, C3H10T12, is induced to undergo adipo-
genesis by a mixture of factors that includes a glucocorticoid such
as dexamethasone. We found that expression of myostatin
(MSTN), a TGF-
family member extensively studied in muscle, was
induced by dexamethasone under those differentiation conditions.
Moreover, MSTN could substitute for dexamethasone in the adi-
pogenesis mixture. However, the adipocytes induced by MSTN in
both cell culture and transgenic mice were small and expressed
markers characteristic of immature adipocytes. These adipocytes
exhibited cell-autonomous increases in insulin sensitivity and glu-
cose oxidation. In mice, these effects produced elevated systemic
insulin sensitivity and resistance to diet-induced obesity. Modula-
tion of the final stages of adipogenesis may provide a novel
approach to understanding and treating metabolic disease.
diabetes obesity
O
besity is associated with a metabolic imbalance that pro-
duces increased adipose tissue mass caused by a combina-
tion of hypertrophy and hyperplasia of adipoc ytes (1). Impor-
t antly, adipose tissue is itself a critical regulator of systemic
met abolism (2), which suggests that modulation of adipogenesis
should have a greater systemic impact than merely altering
adipose tissue mass. Indeed, underst anding the signals involved
in regulating adipogenesis might have broad implications for
diseases such as diabetes.
Adipoc ytes originate from pluripotent progen itor cells that
are recruited to the adipocyte fate (3, 4). A group of factors and
c ytokines, not fully defined, initiate a differentiation cascade in
which a stem cell commits to a lineage, for example, to become
a ‘‘preadipocyte.’’ Further signals continue the progression by
inducing proliferation and finally differentiation into lipid-laden
adipoc ytes (5). To better define these signals, tissue culture
models of adipogenesis have been generated. These studies have
identified a ‘‘mixture’’ of factors that includes dexamethasone,
insulin, and isobut ylmethylxanthine (DIM) that can trigger
adipogenesis in cell culture (5). However, specific roles for each
c omponent of this mixture are poorly understood.
Myost atin (MSTN), also known as growth and differentiation
factor 8 (GDF8), is a member of the TGF-
super family. MSTN
function has been extensively studied in muscle tissue (6, 7),
whereas its role in adipose tissue is not well understood. For
example, dexamethasone induces MSTN expression in muscle
cells (8), but it was not known whether this regulation also occurs
during adipogenesis, or whether MSTN contributes to the dif-
ferentiation process. Consistent with a role for MSTN in early
st ages of adipogenesis, reduction of MSTN expression or activit y
results in decreased adipose tissue in mice (9–11). In addition,
other studies suggested that MSTN plays a role in cell fate
decisions, biasing differentiation toward the adipocyte lineage
and away from the muscle lineage (12). Paradoxically, however,
studies of MSTN expression and action in preadipocytes imply
that MSTN inhibits rather than promotes the formation of
mature adipocy tes (13, 14).
We sought to probe the relationship between MSTN and the
dexamethasone component of the DIM mixture and better
underst and the functions and actions of these signals. We studied
how these signals modulate adipogenesis in cell culture and mice
and evaluated the cellular and systemic consequences of their
actions.
Results
MSTN Is Induced by Dexamethasone in C3H10T12 Cells and Contrib-
utes to Adipogenesis.
We first examined whether MSTN is in-
duced by dexamethasone in the C3H10T12 cell line, which can
dif ferentiate into adipose, muscle, bone, or cartilage under
specific conditions (15); thus, C3H10T1 2 displays characteris-
tics of mesenchymal stem cells. DIM induces adipogenesis in this
cell line (16). The synthetic glucocorticoid dexamethasone is an
essential component of the mixture, as the c ombination of
insulin and isobutylmethylxanthine did not lead to adipogenesis
(Fig. 1). By quantitative real-time PCR, we found that MSTN
ex pression was undetectable in C3H10T12 in the absence of
dexamethasone, and that it was significantly induced (100-fold)
af ter a 4-h exposure to dexamethasone. Thus, MSTN expression
is strongly regulated in C3H10T12 cells by the gluc ocorticoid
c omponent of the differentiation mixture.
We then examined whether adding rec ombinant purified
MSTN could substitute for dexamethasone in the DIM mixture
(MIM). As shown in Fig. 1, the MIM mixture induced significant
levels of adipogenesis as defined by the ac cumulation of intra-
cellular lipid droplets as assessed by Oil-Red-O staining (13).
For comparison, we examined 3T3-L1 cells, a widely studied
cell line that is further differentiated than the C3H10T12 cells
because it is committed to the adipocyte lineage. Although
3T3-L1 has not completed differentiation and therefore is
c ommonly called a preadipocyte cell line, it can be ef ficiently
dif ferentiated to mature adipocytes by the DIM mixture. Unlike
our finding in C3H10T12 cells, the DIM mixture but not the
MIM mixture induced adipogenesis in 3T3-L1 cells (Fig. 1). This
result is consistent with a prior report studying MSTN action in
this cell line (13). Taken at face value, our results suggest that
MSTN can induce adipogenesis as part of the MIM mixture if it
is presented to very-early-stage cells such as mesenchymal stem
cells, but that it has lost its efficacy in adipocyte-c ommitted cells
such as 3T3-L1.
MSTN-Induced Adipocytes Are Small and Apparently Immature. In-
terestingly, the C3H10T12 adipocytes induced by MIM were
smaller than those produced by DIM treatment (Fig. 2). To study
these small adipoc ytes further, we measured the expression of a
series of adipocyte-specific genes, three of which [peroxisome
Author contributions: B.J.F. and K.R.Y. designed research; B.J.F. and R.S.S. performed
research; R.V.F. contributed new reagentsanalytic tools; B.J.F. and K.R.Y. analyzed data;
and B.J.F. and K.R.Y. wrote the paper.
The authors declare no conflict of interest.
Freely available online through the PNAS open access option.
Abbreviations: MSTN, myostatin; DIM, dexamethasoneinsulinisobutylmethylxanthine;
MIM, MSTNinsulinisobutylmethylxanthine; aP2, adipocyte P2.
To whom correspondence should be addressed. E-mail: yamamoto@cmp.ucsf.edu.
© 2006 by The National Academy of Sciences of the USA
www.pnas.orgcgidoi10.1073pnas.0607501103 PNAS
October 17, 2006
vol. 103
no. 42
15675–15680
PHYSIOLOGY
proliferator-activator receptor
, the fatty acid-binding protein
adipoc yte P2 (aP2), and lipoprotein lipase] exhibit elevated
ex pression levels in mature adipocytes, whereas one (preadipo-
c yte factor 1) is expressed at lower levels in mature adipocytes
c ompared with earlier stages in adipogenesis (16, 17). The small
adipoc ytes induced by MSTN expressed these markers in a
pattern similar to that observed before completion of adipogen-
esis (Fig. 3A), suggesting that they may be immature adipoc ytes.
Mature adipose tissue displays endocrine functions, secreting
a number of adipokines (2). Therefore, we measured adipokine
ex pression in the adipoc ytes generated after MSTN exposure,
c ompared with that in adipoc ytes induced by dexamethasone.
The small adipocytes expressed lower levels of adipok ines than
those induced by dexamethasone, further supporting the hypoth-
esis that the MSTN-induced adipocytes are immature (Fig. 3A).
Adipocytes in aP2-MSTN Transgenic Mice. We next investigated the
in vivo relevance of our in vitro findings by generating transgenic
mice that express MSTN in cells undergoing adipogenesis;
specifically, we constructed a MSTN transgene under the c ontrol
of the aP2 promoter and regulatory element (aP2-MSTN). This
promoter is active in the C3H10T12 mesenchymal cell line and
mouse bone marrow and remains active throughout adipogenesis
(Fig. 4A). Use of this promoter and regulatory element has also
been validated in other animal models; for example, aP2-Wnt10b
mice, generated with the same transgene promoter and enhancer
element, display altered mesenchymal stem cell fate (18).
A lthough aP2-MSTN mice were smaller than their WT litter-
mates (Fig. 4C), they otherwise appeared proportional, healthy,
and without gross malformations. However, adipose tissue f rom
the aP2-MSTN transgenic mice displayed a gene expression
profile and reduced cell size similar to the adipocytes differen-
tiated by the MIM mixture in vitro (Fig. 3A and Fig. 6, which is
published as supporting information on the PNAS web site). This
profile was in contrast to the expression pattern found in the
adipose tissue from the WT mice, which, as expected, displayed
the mature adipocyte profile. Consistent with these findings, we
showed that serum levels of the adipokine, adiponectin, were
reduced in the aP2-MSTN mice relative to WT (Fig. 3B).
Therefore, both in vitro and in vivo, MSTN leads to the formation
of adipocy tes with a distinct ex pression profile.
Metabolic Effects of MSTN Transgene Expression in Adipose Tissue.
We next investigated the systemic metabolic c onsequences of the
altered adipose tissue in aP2-MSTN mice. We performed glu-
c ose tolerance tests on lean mice fed regular chow. Compared
with their WT littermates, the transgenic animals exhibited
sign ificantly higher levels of insulin sensitivity, as demonstrated
by their lower fasting glucose levels, lower hyperglycemia, and
more rapid return to euglycemia (Fig. 4B Upper).
We next performed gluc ose tolerance tests on mice fed a
high-fat diet for 7 weeks. WT animals showed evidence of insulin
resistance as demonstrated by hyperglycemia and delayed return
to euglycemia after receiving a glucose bolus. This finding is
c ommon in mice and humans with insulin resist ance and is
c ommonly used as a diagnostic test for type II diabetes in humans
(19). In striking contrast, the aP2-MSTN transgenic mice main-
t ained normal insulin sensitivity on the high-fat diet, as assessed
by their lower fasting blood glucose and normal glucose toler-
ance tests (Fig. 4B Lower).
Adipocyte Autonomous Increase in Insulin Sensitivity with MSTN
Transgene Expression. To determine whether systemic insulin
sensitivity could be attributed to the altered adipocytes, primary
adipoc ytes were purified from WT and transgenic animals, and
insulin sensitivity was tested by measuring upt ake of [
3
H]2-
deoxyglucose after exposure to insulin. Adipocytes harvested
f rom WT an imals responded to insulin with a 3-fold increase in
intracellular gluc ose concentration, whereas adipocytes f rom the
aP2-MSTN animals displayed a 7-fold increase. These results
suggest that the insulin sensitivity of the aP2-MSTN mice is, at
least in part, a cell-autonomous ef fect in adipocytes. To assess
the possibility that a systemic factor (such as obesity) was
af fecting these results, the same experiment was perfor med on
C3H10T1 2 cells that had been differentiated into adipocy tes in
vitro with either MIM or DIM. Adipocytes generated in culture
with MIM were 3-fold more insulin sensitive than DIM-induced
adipoc ytes.
Favorable Systemic Insulin, Glucose, and Triglyceride Levels in aP2-
MSTN Mice. To test whether the immature adipocytes were
protective against other metabolic derangements, we measured
fasting insulin, gluc ose, and triglyceride levels in the aP2-MSTN
mice compared with age-matched WT animals, both maint ained
Fig. 1. MIM induced adipogenesis in C3H10T12 cells but not in 3T3-L1 cells.
Either C3H10T12(Upper) or 3T3-L1 (Lower) cells were grown to confluence.
Cells were then exposed to isobutylmethylxanthine and insulin alone (IM,
Left) or with DIM (Center) or MIM (Right). After 6–10 days of culture, cells were
fixed and stained with Oil-Red-O to identify adipocytes. IM was not suffi-
cient to induce adipogenesis in either cell line. DIM induced adipogenesis
in both cell lines. In C3H10T12 cells, but not in 3T3-L1 cells, MIM induced
adipogenesis.
Fig. 2. MSTN induced adipogenesis in C3H10T12 cells. Adipogenesis was
induced in confluent C3H10T12 cells with either MIM (A, C, and E) or DIM (B,
D, and F). Cells were photographed under light microscopy while in culture (A
and B) or after fixation and staining with Oil-Red-O (CF). Adipocytes formed
after MSTN treatment were smaller and appeared to contain less lipid than the
adipocytes formed after dexamethasone treatment. Cells exposed to either
isobutylmethylxanthine and insulin or each component alone had minimal
adipogenesis (data not shown). (Magnification: A, B, E, and F, 200; C and D,
100.)
15676
www.pnas.orgcgidoi10.1073pnas.0607501103 Feldman et al.
on a high-fat diet. The transgenic animals did not display any of
the characteristics of metabolic syndrome that were evident in
the WT mice. In particular, the transgenic mice had significantly
lower fasting insulin, glucose, and triglyceride levels than WT
mice (Fig. 4D).
aP2-MSTN Transgenic Mice Have an Increased Metabolic Rate and Are
Resistant to Diet-Induced Obesity. Energy balance was examined in
lean (chow-fed) transgenic and WT littermate mice in metabolic
cages, which continuously measure food intake, locomotor ac-
tivity, and metabolic rate as calculated by oxygen consumption
(VO
2
). aP2-MSTN mice had a higher metabolic rate than WT
mice (P 0.005), whereas activit y and food intake (normalized
to total body mass) were similar (Fig. 5A). This finding suggested
that the transgenic animals might be resistant to diet-induced
obesity. Consequently, transgenic and c ontrol animals were
placed on the high-fat diet for 7 weeks, and the quantity of
adipose tissue was monitored by dual energy x-ray absorptiom-
etry scanning. As ex pected, WT mice became obese on this diet
with a mean body fat of 38%, whereas the aP2-MSTN mice were
resistant to weight gain, reaching a maximum mean body fat of
26% (Fig. 4C). Therefore, the adipocytes present in the aP2-
MSTN animals appear to produce a higher met abolic rate and
resistance to the development of obesity.
MSTN Transgene Expression Leads to Increased Glucose Oxidation.
We next studied oxygen consumption and aerobic metabolism in
lean animals, using met abolic cages to measure the respiratory
exchange ratio (VCO
2
VO
2
). Although not statistically signifi-
cant, the ratio was greater in transgenic animals than in WT
an imals (Fig. 5A), which suggested that the aP2-MSTN mice may
oxidize glucose to generate energy to a greater extent than the
WT mice. Consistent with this view, quantitative real-time PCR
revealed increased ex pression of mRNA-enc oding glucose trans-
porters GLUT1 and GLUT4, and critical glycoly tic enzymes
hexok inase, glucokinase, pyruvate k inase, and pyr uvate dehy-
drogenase, in adipose tissue from aP2-MSTN mice. Further-
more, mRNAs for fat biosynthesis enzy mes such as fatty acid
synthase and acyl CoA:diacylglycerol acyltransferase 2 were
down-regulated in the transgenic adipose tissue compared with
WT (Fig. 5B). These results suggest that the increased metabolic
rate and resistance to obesity found in the transgenic an imals
reflects increased glucose uptake and glucose oxidation in the
adipose tissue.
To determine whether MSTN-induced adipocytes use more
gluc ose in a cell autonomous manner, we directly measured
gluc ose oxidation, using
14
C-labeled glucose (see Methods), in
C3H10T1 2 cells that had been differentiated to adipocytes with
MIM exposure compared with adipoc ytes generated by DIM
treatment. As predicted, the adipocytes generated after MSTN
treatment had sign ificantly (P 0.001) more glucose oxidation
than dexamethasone-induced adipocytes (Fig. 5C).
Discussion
We have demonstrated that MSTN is induced by dexamethasone
in mesenchymal stem cells, and that it can substitute for dexa-
methasone to induce adipogenesis. In c ontrast, MSTN fails to
trigger adipogenesis in the 3T3-L1 preadipocyte cell line. We
speculate that a sensitive interval may exist during cell fate
deter mination and early differentiation of mesenchymal stem
cells in which MSTN can induce adipogenesis, and that the
sensitive period ends before the preadipocyte stage represented
Fig. 3. Adipocytes generated with MSTN exposure have the expression profile of an immature adipocyte. (A) Real-time quantitative PCR was used to compare
the expression patterns between adipocytes formed in culture after MIM or DIM exposure (n 4 for each condition). (Upper Left) MSTN-induced adipocytes had
lower levels of mature adipocyte markers peroxisome proliferator-activator receptor
(PPAR), aP2, and lipoprotein lipase (LPL) and a higher level of the immature
marker preadipocyte factor 1 (Pref-1). (Upper Right) Cultured MSTN-induced adipocytes also had lower expression levels of adipokines compared with
dexamethasone-induced adipocytes. (Lower) MSTN transgenic (TG) mice had an analogues immature adipose tissue expression profile and lower levels of
adipokine expression. (B) Transgenic animals had lower levels of circulating adiponectin as measured by ELISA performed on serum from transgenic (Tg) mice
and WT littermates. Error bars represent SDs.
Feldman et al. PNAS
October 17, 2006
vol. 103
no. 42
15677
PHYSIOLOGY
by the 3T3-L1 cells. This hypothesis could explain the seemingly
c onflicting effects of MSTN reported previously. In addition,
this hypothesis is c onsistent with the findings in the MSTN
k nockout mice. These animals have decreased adipose tissue,
which may result from impaired initiation of adipogenesis.
Import antly, adipogenesis induced in C3H10T12 by MSTN
yielded cells that maint ain the ex pression profile of immature
adipoc ytes. Furthermore, expression of MSTN during adipogen-
esis in vivo produced similar results, with the adipose tissue in
these mice maintaining this distinct expression pattern. We
postulate that the adipoc ytes generated after MSTN exposure
represent a novel stage in adipogenesis that is further differen-
tiated than the preadipoc yte but earlier than the fully differen-
tiated adipocyte. According to this model, dexamethasone-
c ontaining dif ferentiation conditions such as the DIM mixture
would trigger cells to pass through this stage, whereas MSTN-
c ontaining conditions such as the MIM mixture would lead to
ac cumulation of the immature adipocytes. Our findings in the
MSTN transgen ic mice suggest that MSTN expression may
dominantly inhibit this later stage in adipogenesis, as the an imals
are, of course, glucocorticoid replete. Indeed, we have observed
such dominant negative behavior of MSTN in C3H10T12 cells
(B.J.F., unpublished results). In future studies, it will be inter-
esting to examine the mechanistic relationships of the glucoc or-
tic oid and MSTN signals in adipogenesis and determine whether
the MSTN-induced cells are precursors along the full differen-
tiation pathway, or rather represent a branch that exits from the
pathway.
Finally, it is notable that the MSTN-induced adipocytes have
a favorable impact on metabolism. In particular, metabolic
studies revealed that these changes result in improved insulin
sensitivity and resistance to obesity, which results, at least in part,
f rom increased glucose oxidation. These findings suggest that
altering the final st ages of adipogenesis, for example, through
Fig. 4. aP2-MSTN mice have a favorable metabolic profile and are resistant to obesity. (A) The transgene (Tg) construct was created by subcloning the full-length
MSTN cDNA behind an aP2 promoter. Transgenic mice had a 10-fold overexpression of MSTN as detected by quantitative real-time PCR amplification of cDNA
generated from the adipose tissue of WT and transgenic mice. Primers P1 and P2 were used to amplify a transgene-specific cDNA, whereas primers P1 and P3
were used to compare total MSTN expression levels. Quantitative PCR amplification of hypoxanthine-guanine phosphoribosyl transferase (HPRT) was used to
control for cDNA concentration between samples. (B) Glucose tolerance tests were performed on both lean mice (Upper) and mice on a high-fat diet (Lower).
Blood sugar levels were checked after an overnight fast (time 0) and then 2 gkg of glucose was injected i.p. After the glucose bolus, tail blood sugar levels
were checked at each time point. At the 60-min time point, all WT animals on the high-fat diet had blood sugars above the upper limit of the meter, therefore
the upper limit value (600 mgdl) was used at that time point. aP2-MSTN mice were more insulin sensitive than their WT littermates on both regular (P 0.0061)
and high-fat chow (P 0.0025). (C) aP2-MSTN mice and their WT littermates were placed on a high-fat diet for 7 weeks. Transgenic mice (Lower Right) were
resistant to the obesity that developed in WT animals (Lower Left). (D) aP2-MSTN mice had lower fasting glucose, insulin, and triglyceride levels than their WT
littermates. Error bars represent SDs.
15678
www.pnas.orgcgidoi10.1073pnas.0607501103 Feldman et al.
t argeted MSTN ex posure, may have a therapeutic benefit for
humans with metabolic diseases such as diabetes.
Methods
Tissue Culture and Differentiation Assays. C3H10T12 cells were
obt ained from the Un iversity of California, San Francisco tissue
culture core facility. 3T3-L1 cells were obtained from ATCC
(Manassas, VA). Cells were grown to confluence in high-gluc ose
DMEM supplemented with 1 mM pyr uvate and 10% FBS. Two
days af ter c onfluence cells were exposed to either 1
M dexa-
methasone (Sigma, St. Louis, MO) or 40 nM recombinant MSTN
(R&D Systems, Minneapolis, MN) in addition to 11.5
gml
isobut ylmethylxanthine (Sigma) and 24 ngml insulin (Sigma).
Cells were cultured for 10 days, and half of the media was
replaced with fresh DMEM and 10% FBS every 2 days.
Quantitative PCR. Quantitative PCR was perfor med as described
(20) with an Opticon2 machine and sof tware (MJ Research,
Cambridge, MA). Primer sequences used are available on re-
quest.
Generation of Transgenic Mice. MSTN cDNA was amplified from
mouse muscle RNA (Stratagene, La Jolla, CA). The cDNA was
subcloned behind the aP2 promoter and regulatory domain and
in front of a poly(A) sequence (gifts from Ormond MacDougald,
Un iversity of Michigan, A nn Arbor, MI). Transgene microin-
jection and transplant into pseudopregnant mice was performed
by the Un iversity of California, San Francisco transgenic cancer
c ore facility using the FVB strain. Genot yping was done by PCR
using transgene-specific primers. A ll mice were housed in a
pathogen-f ree barrier-type facility (12-h light12-h dark cycle).
Male mice were used for all studies, and they were fed either a
st andard chow diet (5053 PicoLab Diet; Purina, St. Louis, MO)
or a high-fat, Western-type diet (TD.01064, Harlan-Teklad,
Madison, WI) that c ontains 20% anhydrous milk fat, 1% corn
oil, and 0.2% cholesterol by weight. All an imal studies were
approved by the Un iversity of Californ ia, San Francisco Com-
mittee on Animal Research.
Glucose Tolerance Testing. Male transgenic (n 5) and WT (n
7) littermate animals, at 3–4 months of age, were fasted over-
n ight. Fasting and subsequent glucose levels were obtained from
t ail vein blood with a One Touch Ultra glucometer and test
strips. Glucose (2 gkg) dissolved in sterile saline was injected
into the peritoneal cavit y of the mice, and blood glucose was
mon itored.
Isolation of Primary Adipocytes and Determination of Glucose Trans-
port. Insulin-stimulated glucose transport in purified primary
white adipocytes from transgenic and WT littermates was de-
ter mined as described (1, 21). Briefly, fat depots where treated
with t ype I collagenase (2 mgml; Worthington Biochemical,
L akewood, NJ) and filtered through 500-
m nylon mesh. Adi-
poc ytes were washed three times with Krebs-Ringer-Hepes
buf fer and 2.5% BSA. Cells were incubated with [
3
H]2-
deoxyglucose with and without insulin for 30 min. Adipocytes
Fig. 5. aP2-MSTN transgenic mice have an increased metabolic rate with increased glycolysis in their adipose tissue compared with WT littermates. (A)
Transgenic (Tg) and WT littermates were placed in metabolic cages that simultaneously measures metabolic rate (as calculated by gas exchange), activity, food
intake, and weight. Transgenic animals had an increased metabolic rate (P 0.005), slightly increased food intake, and similar activity as WT littermates. This
increased metabolic rate likely contributes to protecting the animals from obesity. (B) Quantitative real-time PCR was used to measure the expression levels of
genes involved in glucose utilization in the adipose tissue of the mice. Transgenic animals had higher expression levels than WT littermates of genes involved
in glucose transport (GLUT1 and GLUT4) and enzymes in the glycolysis pathway [glucokinase (GK), hexokinase (HK), and pyruvate kinase (PK)]. In addition, the
aP2-MSTN animals had lower expression levels of some of the genes involved in lipogenesis [fatty acid synthase (FAS) and diacylglycerol acyltransferase 2
(DGAT2)]. (C) The rate of glucose oxidation in adipocytes differentiated in tissue culture was directly measured by using
14
C glucose (see Methods). Adipocytes
generated after exposure to MIM have a cell autonomous higher rate of glucose oxidation than adipocytes generated after DIM exposure (P 0.001). Error bars
represent SDs.
Feldman et al. PNAS
October 17, 2006
vol. 103
no. 42
15679
PHYSIOLOGY
were purified from media by passing through silic on oil DC550
with phthalic acid dinonyl ester (ratio 2:3). Intracellular [
3
H]2-
deoxyglucose was quantified with a scintillation counter. Adi-
poc yte yield was variable between samples, therefore, insulin
sensitivity was internally normalized by calculating fold change
bet ween basal and insulin-induced samples.
Rate of Glucose Oxidation. Rate of glucose oxidation was deter-
mined as described (22). Briefly, confluent C3H10T12 cells
were treated with MIM, DIM, or vehicle and cultured for 10 days
as described. Cells were washed and st arved for2hinKrebes-
Ringer-Hepes buffer. C
14
gluc ose alone (basal) or with insulin
was added to the media and cells were incubated at 37°C for 2 h.
14
CO
2
was released with the addition of 2 M HCL and quantified
by scintillation counting.
Body Composition. Mice were fasted for 4 h and anesthetized with
isoflurane, and their body composition was analyzed by dual
energy x-ray absorptiometry with a PixiMus2 scanner (GE
Healthcare Lunar, Madison, WI).
Energy Balance. Food intake and oxygen consumption (VO
2
)
were measured by indirect calorimetry (Oxymax Comprehensive
L ab An imal Monitoring System, Columbus Instruments, Co-
lumbus, OH) over 3 days. Both parameters were normalized to
lean body mass, as measured by dual energy x-ray absorptiom-
etry scanning on the day of initiating calorimetry studies. Studies
were performed on three WT and five transgen ic male mice. P
values were calculated by using two-tailed t tests.
We thank Ormand MacDougald for providing the aP2 promoter and
enhancer construct; Wally Wang, Stefan Taubert, and David Feldman
for comments on the manuscript; members of K.R.Y.’s laboratory for
valuable discussions; and Valerie Doughert y for administrative support.
This work was supported by National Institutes of Health Grants
CA20535 (to K.R.Y.), DK56084 (to R.V.F.), DK07161 and DK73697 (to
B.J.F.), and DK56084 (to R.S.S.).
1. Shepherd PR, Gnudi L, Tozzo E, Yang H, Leach F, Kahn BB (1993) J Biol
Chem 268:22243–22246.
2. Kershaw EE, Flier JS (2004) J Clin Endocrinol Metab 89:2548–2556.
3. Yu ZK, Wright JT, Hausman GJ (1997) Obes Res 5:9–15.
4. Tang QQ, Otto TC, Lane MD (2004) Proc Natl Acad Sci USA 101:9607–
9611.
5. MacDougald OA, Mandrup S (2002) Trends Endocrinol Metab 13:5–11.
6. McPherron AC, Lawler AM, Lee SJ (1997) Nature 387:83–90.
7. McNally EM (2004) N Engl J Med 350:2642–2644.
8. Ma K, Mallidis C, Artaza J, Taylor W, Gonzalez-Cadav id N, Bhasin S (2001)
Am J Physiol 281:E1128–E1136.
9. McPherron AC, Lee SJ (2002) J Clin Invest 109:595–601.
10. Lin J, Arnold HB, Della-Fera M A, Azain MJ, Hartzell DL, Baile CA (2002)
Biochem Biophys Res Commun 291:701–706.
11. Zhao B, Wall RJ, Yang J (2005) Biochem Biophys Res Commun 337:248–255.
12. Artaza JN, Bhasin S, Magee TR, Reisz-Porszasz S, Shen R, Groome NP, Fareez
MM, Gonzalez-Cadavid NF (2005) Endocrinology 146:3547–3557.
13. Kim HS, Liang L, Dean RG, Hausman DB, Hartzell DL, Baile CA (2001)
Biochem Biophys Res Commun 281:902–906.
14. Hirai S, Matsumoto H, Hino N, Kawachi H, Matsui T, Yano H (2006) Domest
Anim Endocrinol, in press.
15. Taylor SM, Jones PA (1979) Cell 17:771–779.
16. Rosen ED, Spiegelman BM (2000) Annu Rev Cell Dev Biol 16:145–171.
17. Smas CM, Sul HS (1993) Cell 73:725–734.
18. Bennett CN, Longo KA, Wright WS, Suva LJ, Lane TF, Hankenson KD,
MacDougald OA (2005) Proc Natl Acad Sci USA 102:3324–3329.
19. American Diabetes Association (2006) Diabetes Care 29(Suppl 1):S43–S48.
20. Taubert S, Van Gilst MR, Hansen M, Yamamoto KR (2006) Genes Dev
20:1137–1149.
21. Konrad D, Bilan PJ, Nawaz Z, Sweeney G, Niu W, Liu Z, Antonescu CN,
Rudich A, Klip A (2002) Diabetes 51:2719–2726.
22. Powelka AM, Seth A, Virbasius JV, Kiskinis E, Nicoloro SM, Guilherme A,
Tang X, Straubhaar J, Cherniack AD, Parker MG, Czech MP (2006) J Clin
Invest 116:125–136.
15680
www.pnas.orgcgidoi10.1073pnas.0607501103 Feldman et al.
... MSTN, a member of the TGF-b superfamily, is a secreted protein that is specifically expressed in skeletal muscle, and is associated with myogenesis and adipogenesis in muscle development and regeneration (76,77 (76). Additionally, Feldman and colleagues showed that MSTN can serve as a substitute for dexamethasone in inducing adipogenesis in C3H10T (1/2) cells but not in 3T3-L1 preadipocytes, which indicates that MSTN plays a role in promoting adipogenesis in the specific early stage (77). ...
... MSTN, a member of the TGF-b superfamily, is a secreted protein that is specifically expressed in skeletal muscle, and is associated with myogenesis and adipogenesis in muscle development and regeneration (76,77 (76). Additionally, Feldman and colleagues showed that MSTN can serve as a substitute for dexamethasone in inducing adipogenesis in C3H10T (1/2) cells but not in 3T3-L1 preadipocytes, which indicates that MSTN plays a role in promoting adipogenesis in the specific early stage (77). It should be noted that the adipocytes induced by MSTN in cell cultures and transgenic mice revealed the expression of markers associated with immature adipocytes, which exhibit favorable metabolic effects (77). ...
... Additionally, Feldman and colleagues showed that MSTN can serve as a substitute for dexamethasone in inducing adipogenesis in C3H10T (1/2) cells but not in 3T3-L1 preadipocytes, which indicates that MSTN plays a role in promoting adipogenesis in the specific early stage (77). It should be noted that the adipocytes induced by MSTN in cell cultures and transgenic mice revealed the expression of markers associated with immature adipocytes, which exhibit favorable metabolic effects (77). However, inconsistent with previous findings, Liu et al. reported that the activated myostatin/SMAD4 signal promotes the expression of miR-124-3p, and inhibits adipogenesis by downregulating the expression of glucocorticoid receptor (GR) in porcine preadipocytes (80). ...
Article
Full-text available
Intermuscular adipose tissue (IMAT) is a unique adipose depot interspersed between muscle fibers (myofibers) or muscle groups. Numerous studies have shown that IMAT is strongly associated with insulin resistance and muscular dysfunction in people with metabolic disease, such as obesity and type 2 diabetes. Moreover, IMAT aggravates obesity-related muscle metabolism disorders via secretory factors. Interestingly, researchers have discovered that intermuscular brown adipocytes in rodent models provide new hope for obesity treatment by acting on energy dissipation, which inspired researchers to explore the underlying regulation of IMAT formation. However, the molecular and cellular properties and regulatory processes of IMAT remain debated. Previous studies have suggested that muscle-derived stem/progenitor cells and other adipose tissue progenitors contribute to the development of IMAT. Adipocytes within IMAT exhibit features that are similar to either white adipocytes or uncoupling protein 1 (UCP1)-positive brown adipocytes. Additionally, given the heterogeneity of skeletal muscle, which comprises myofibers, satellite cells, and resident mesenchymal progenitors, it is plausible that interplay between these cellular components actively participate in the regulation of intermuscular adipogenesis. In this context, we review recent studies associated with IMAT to offer insights into the cellular origins, biological properties, and regulatory mechanisms of IMAT. Our aim is to provide novel ideas for the therapeutic strategy of IMAT and the development of new drugs targeting IMAT-related metabolic diseases.
... The precise role of myostatin in adipocyte differentiation is controversial with studies indicating both pro-and anti-adipogenic effects of myostatin. In committed adipocyte cell lines (e.g., 3T3-L1 pre-adipocytes) myostatin inhibits cell differentiation [27], whereas in pluripotent stem cells (e.g., CH3 10T1/2, a mesenchymal progenitor) myostatin contributes to differentiation towards an adipogenic phenotype suggesting that it plays a role in the early commitment to the adipocyte cell lineage and away from the myocyte lineage [28,29]. Consistent with this suggestion, myostatin-null animals have markedly reduced fat mass alongside their increase in muscle mass [30]. ...
... Although our data cannot demonstrate causality, they raise the possibility that the accumulation of extramuscular limb fat may be influenced by activin receptor signalling-a hypothesis consistent with known effects of both myostatin and activin. Previous studies have shown that activin A signalling increases the proliferation of adipocyte progenitors but reduces adipocyte differentiation [40,41] and myostatin increases the commitment of pluripotent mesenchymal cells to the adipocyte lineage and away from the myogenic lineage [28,29]. By increasing the pool of adipocyte progenitors therefore, increased activin signalling may increase fat mass over time in an analogous manner to the effect of bone morphogenetic proteins (BMPs) on muscle formation and hypertrophy. ...
Article
Full-text available
Background Ageing is associated with changes in body composition including an overall reduction in muscle mass and a proportionate increase in fat mass. Sarcopenia is characterised by losses in both muscle mass and strength. Body composition and muscle strength are at least in part genetically determined, consequently polymorphisms in pathways important in muscle biology (e.g., the activin/myostatin signalling pathway) are hypothesised to contribute to the development of sarcopenia. Methods We compared regional body composition measured by DXA with genotypes for two polymorphisms (rs10783486, minor allele frequency (MAF) = 0.26 and rs2854464, MAF = 0.26) in the activin 1B receptor ( ACVR1B ) determined by PCR in a cross-sectional analysis of DNA from 110 older individuals with sarcopenia from the LACE trial. Results Neither muscle mass nor strength showed any significant associations with either genotype in this cohort. Initial analysis of rs10783486 showed that males with the AA/AG genotype were taller than GG males (174±7cm vs 170±5cm, p = 0.023) and had higher arm fat mass, (median higher by 15%, p = 0.008), and leg fat mass (median higher by 14%, p = 0.042). After correcting for height, arm fat mass remained significantly higher (median higher by 4% p adj = 0.024). No associations (adjusted or unadjusted) were seen in females. Similar analysis of the rs2854464 allele showed a similar pattern with the presence of the minor allele (GG/AG) being associated with greater height (GG/AG = 174±7 cm vs AA = 170 ±5cm, p = 0.017) and greater arm fat mass (median higher by 16%, p = 0.023). Again, the difference in arm fat remained after correction for height. No similar associations were seen in females analysed alone. Conclusion These data suggest that polymorphic variation in the ACVR1B locus could be associated with body composition in older males. The activin/myostatin pathway might offer a novel potential target to prevent fat accumulation in older individuals.
... Myostatin is a myokine involved in the maintenance of metabolic homeostasis and the regulation of adipose tissue function. 10,62 Acute exercise increases the concentration of myostatin in serum in humans 63,S25 ; however, another study showed a decrease in the concentration of myostatin in serum in healthy young men 24 h after resistance, 64 which may be attributed to the different time points of blood sampling. Mutation or knockout of myostatin results in muscle hypertrophy in humans and mice. ...
Article
Full-text available
Exercise is recognized to play an observable role in improving human health, especially in promoting muscle hypertrophy and intervening in muscle mass loss‐related diseases, including sarcopenia. Recent rapid advances have demonstrated that exercise induces the release of abundant cytokines from several tissues (e.g., liver, muscle, and adipose tissue), and multiple cytokines improve the functions or expand the numbers of adult stem cells, providing candidate cytokines for alleviating a wide range of diseases. Muscle satellite cells (SCs) are a population of muscle stem cells that are mitotically quiescent but exit from the dormancy state to become activated in response to physical stimuli, after which SCs undergo asymmetric divisions to generate new SCs (stem cell pool maintenance) and commit to later differentiation into myocytes (skeletal muscle replenishment). SCs are essential for the postnatal growth, maintenance, and regeneration of skeletal muscle. Emerging evidence reveals that exercise regulates muscle function largely via the exercise‐induced cytokines that govern SC potential, but this phenomenon is complicated and confusing. This review provides a comprehensive integrative overview of the identified exercise‐induced cytokines and the roles of these cytokines in SC function, providing a more complete picture regarding the mechanism of SC homeostasis and rejuvenation therapies for skeletal muscle.
... Interestingly, we found that activation of PPARγ-mediated adipocyte hyperplasia partially reduced high-fat diet-induced excessive accumulation of adipose tissue. Studies have shown that activation of adipocyte hyperplasia could increase the metabolic rate of adipose tissue (Feldman et al., 2006). In this study, the (dgat1a, dgat1b, dgat2, atgl, cpt1b, lpl) in liver. ...
... Snail1 plays a key role in TGF-β1-induced maintenance of stemness in MSCs [89]. Myogenesis inhibitor is a key negative regulator of skeletal mass, promotes adipogenesis, and inhibits myogenesis in C3H10T1/2 pluripotent MSCs [90,91]. Blocking the Notch signaling pathway promoted autophagy-mediated lipogenic differentiation of MSCs via PTEN-PI3K/AKT/mTOR pathway [92]. ...
Article
Full-text available
Mesenchymal stem cells (MSCs), as a kind of pluripotent stem cells, have attracted much attention in orthopedic diseases, geriatric diseases, metabolic diseases, and sports functions due to their osteogenic potential, chondrogenic differentiation ability, and adipocyte differentiation. Anti-inflammation, anti-fibrosis, angiogenesis promotion, neurogenesis, immune regulation, and secreted growth factors, proteases, hormones, cytokines, and chemokines of MSCs have been widely studied in liver and kidney diseases, cardiovascular and cerebrovascular diseases. In recent years, many studies have shown that the extracellular vesicles of MSCs have similar functions to MSCs transplantation in all the above aspects. Here we review the research progress of MSCs and their exocrine vesicles in recent years.
... After the MSTN knockout, the activity of the respiratory chain complex is reduced, and the metabolism of OXPHOS in the muscle is switched to glycolytic metabolism [43]. A knockdown of MSTN up-regulates the expression of glucose transporters HK and PK and accelerates glycolysis [44]. It has also been shown that an MSTN deletion increases creatine kinase (CK) activity [45]. ...
Article
Full-text available
Myostatin (MSTN) is an important negative regulator of skeletal muscle growth in animals. A lack of MSTN promotes lipolysis and glucose metabolism but inhibits oxidative phosphorylation (OXPHOS). Here, we aimed to investigate the possible mechanism of MSTN regulating the mitochondrial energy homeostasis of skeletal muscle. To this end, MSTN knockout mice were generated by the CRISPR/Cas9 technique. Expectedly, the MSTN null (Mstn−/−) mouse has a hypermuscular phenotype. The muscle metabolism of the Mstn−/− mice was detected by an enzyme-linked immunosorbent assay, indirect calorimetry, ChIP-qPCR, and RT-qPCR. The resting metabolic rate and body temperature of the Mstn−/− mice were significantly reduced. The loss of MSTN not only significantly inhibited the production of ATP by OXPHOS and decreased the activity of respiratory chain complexes, but also inhibited key rate-limiting enzymes related to the TCA cycle and significantly reduced the ratio of NADH/NAD+ in the Mstn−/− mice, which then greatly reduced the total amount of ATP. Further ChIP-qPCR results confirmed that the lack of MSTN inhibited both the TCA cycle and OXPHOS, resulting in decreased ATP production. The reason may be that Smad2/3 is not sufficiently bound to the promoter region of the rate-limiting enzymes Idh2 and Idh3a of the TCA cycle, thus affecting their transcription.
Article
Skeletal muscle plays a vital role in the regulation of systemic metabolism, partly through its secretion of endocrine factors which are collectively known as myokines. Altered myokine levels are associated with metabolic diseases, such as type 2 diabetes (T2D). The significance of interorgan crosstalk, particularly through myokines, has emerged as a fundamental aspect of nutrient and energy homeostasis. However, a comprehensive understanding of myokine biology in the setting of obesity and T2D remains a major challenge. In this review, we discuss the regulation and biological functions of key myokines that have been extensively studied during the past two decades, namely interleukin 6 (IL-6), irisin, myostatin (MSTN), growth differentiation factor 11 (GDF11), fibroblast growth factor 21 (FGF21), apelin, brain-derived neurotrophic factor (BDNF), meteorin-like (Metrnl), secreted protein acidic and rich in cysteine (SPARC), β-aminoisobutyric acid (BAIBA), Musclin, and Dickkopf-3 (Dkk3). Related to these, we detail the role of exercise in myokine expression and secretion together with their contributions to metabolic physiology and disease. Despite significant advancements in myokine research, many myokines remain challenging to measure accurately and investigate thoroughly. Hence, new research techniques and detection methods should be developed and rigorously tested. Therefore, developing a comprehensive perspective on myokine biology is crucial, as this will likely offer new insights into the pathophysiological mechanisms underlying obesity and T2D and may reveal novel targets for therapeutic interventions.
Article
Full-text available
The ability to generate thermogenic fat could be a targeted therapy to thwart obesity and improve metabolic health. Brown and beige adipocytes are two types of thermogenic fat cells that regulate energy balance. Both adipocytes share common morphological, biochemical, and thermogenic properties. Yet, recent evidence suggests unique features exist between brown and beige adipocytes, such as their cellular origin and thermogenic regulatory processes. Beige adipocytes also appear highly plastic, responding to environmental stimuli and interconverting between beige and white adipocyte states. Additionally, beige adipocytes appear to be metabolically heterogenic and have substrate specificity. Nevertheless, obese and aged individuals cannot develop beige adipocytes in response to thermogenic fat-inducers, creating a key clinical hurdle to their therapeutic promise. Thus, elucidating the underlying developmental, molecular, and functional mechanisms that govern thermogenic fat cells will improve our understanding of systemic energy regulation and strive for new targeted therapies to generate thermogenic fat. This review will examine the recent advances in thermogenic fat biogenesis, molecular regulation, and the potential mechanisms for their failure.
Article
Full-text available
To gain insight into the molecular pathogenesis of obesity and specifically the role of nutrient partitioning in the development of obesity, we overexpressed the insulin-responsive glucose transporter (GLUT4) in transgenic mice under the control of the fat-specific aP2 fatty acid-binding protein promoter/enhancer. Two lines of transgenic mice were generated, which overexpressed GLUT4 6-9-fold in white fat and 3-5-fold in brown fat with no overexpression in other tissues. In vivo glucose tolerance was enhanced in transgenic mice. In isolated epididymal, parametrial, and subcutaneous adipose cells from transgenic mice, basal glucose transport was 20-34-fold greater than in nontransgenic littermates. Insulin-stimulated glucose transport was 2-4-fold greater in cells from transgenic mice. Total body lipid was increased 2-3-fold in transgenic mice overexpressing GLUT4 in fat. Surprisingly, fat cell size was unaltered and fat cell number was increased > 2-fold. This is the first animal model in which increased fat mass results solely from adipocyte hyperplasia and it will be a valuable model for understanding the mechanisms responsible for fat cell replication and/or differentiation in vivo.
Article
Full-text available
We cloned and characterized a 3.3-kb fragment containing the 5'-regulatory region of the human myostatin gene. The promoter sequence contains putative muscle growth response elements for glucocorticoid, androgen, thyroid hormone, myogenic differentiation factor 1, myocyte enhancer factor 2, peroxisome proliferator-activated receptor, and nuclear factor-kappaB. To identify sites important for myostatin's gene transcription and regulation, eight deletion constructs were placed in C(2)C(12) and L6 skeletal muscle cells. Transcriptional activity of the constructs was found to be significantly higher in myotubes compared with that of myoblasts. To investigate whether glucocorticoids regulate myostatin gene expression, we incubated both cell lines with dexamethasone. On both occasions, dexamethasone dose dependently increased both the promoter's transcriptional activity and the endogenous myostatin expression. The effects of dexamethasone were blocked when the cells were coincubated with the glucocorticoid receptor antagonist RU-486. These findings suggest that glucocorticoids upregulate myostatin expression by inducing gene transcription, possibly through a glucocorticoid receptor-mediated pathway. We speculate that glucocorticoid-associated muscle atrophy might be due in part to the upregulation of myostatin expression.
Article
Full-text available
The balance of contradictory signals experienced by preadipocytes influences whether these cells undergo adipogenesis. In addition to the endocrine system, these signals originate from the preadipocytes themselves or operate as part of a feedback loop involving mature adipocytes. The factors that regulate adipogenesis either promote or block the cascade of transcription factors that coordinate the differentiation process. Some of the positive factors reviewed include insulin-like growth factor I, macrophage colony-stimulating factor, fatty acids, prostaglandins and glucocorticoids, and negative factors reviewed include Wnt, transforming growth factor beta, inflammatory cytokines and prostaglandin F(2alpha). Tipping the scales towards or away from adipogenesis has profound implications for human health. In this review, we describe recent contributions to the field and will focus on factors that probably play a role in vivo.
Article
Three new mesenchymal phenotypes were expressed in cultures of Swiss 3T3 and C3H/10T1/2CL8 mouse cells treated with 5-azacytidine or 5-aza-2'-deoxycytidine. These phenotypes were characterized as contractile striated muscle cells, biochemically differentiated adipocytes and chondrocytes capable of the biosynthesis of cartilage-specific proteins. The number of muscle and fat cells which appeared in treated cultures was dependent upon the concentration of 5-azacytidine used, but the chondrocyte phenotype was not expressed frequently enough for quantitation. The differentiated cell types were only observed several days or weeks after treatment with the analog, implying that cell division was obligatory for the expression of the new phenotypes. Oncogenically transformed C3H/10T1/2CL8 cells also developed muscle cells after exposure to 5-azacytidine, but at a reduced rate when compared to the parent line. Five subclones of the 10T1/2 line which were the progeny of single cells all expressed both the muscle and fat phenotypes following 5-azacytidine treatment. The effects of the analog are therefore not due to the selection of preexisting myoblasts or adipocytes in the cell populations. Rather, it is possible that 5-azacytidine, after incorporation into DNA, causes a reversion to a more pluripotential state from which the new phenotypes subsequently differentiate.
Article
With the aim of identifying novel regulators of adipocyte differentiation, we have cloned and characterized preadipocyte factor 1 (pref-1), a novel member of the epidermal growth factor (EGF)-like family of proteins. Pref-1 is synthesized as a transmembrane protein with six tandem EGF-like repeats. In preadipocytes, multiple discrete forms of pref-1 protein of 45-60 kd are present, owing in part to N-linked glycosylation. While pref-1 mRNA is abundant in preadipocytes, its expression is completely abolished during differentiation of 3T3-L1 preadipocytes to adipocytes. Moreover, constitutive expression of pref-1 in preadipocytes, which in effect blocks its down-regulation, drastically inhibits adipose differentiation. This indicates that pref-1 functions as a negative regulator of adipocyte differentiation, possibly in a manner analogous to EGF-like proteins that govern cell fate decisions in invertebrates.
Article
Glucocorticoids or the glucocorticoid analog dexamethasone (DEX) enhances the differentiation of preadipocytes in the presence of insulin and influences preadipocyte proliferation. The purpose of the present study was to determine if DEX can induce the recruitment of preadipocytes. Using monoclonal antibodies for complement-mediated cytotoxicity, preadipocytes were removed from porcine stromal vascular (S-V) cell cultures. Our experiments demonstrated for the first time that after removal of preadipocytes by cytotoxicity, preadipocytes or fat cells could be induced by DEX or DEX plus insulin but not by insulin alone. However, many more fat cells were induced (258 +/- 15/unit area) when DEX was added with fetal bovine serum (FBS) followed with insulin treatment, compared to DEX with insulin (21.3 +/- 5.1/ unit area) after removal of preadipocytes. Immunocytochemistry with AD-3, a preadipocyte marker, showed that DEX with FBS for 3 days after seeding (i.e., the proliferation phase) produced many more preadipocytes (AD-3 positive, 223 +/- 45/unit area) than FBS alone (10.5 +/- 1.4/unit area). Bromodeoxyuridine (BrdU) incorporation assays demonstrated that the efficiency of DEX with FBS (i.e., during proliferation) was mitosis dependent. Accordingly, we conclude that: porcine S-V cultures contain preadipocytes at different stages of differentiation; and that DEX induced early preadipocyte differentiation depends on mitosis.
Article
The transforming growth factor-beta (TGF-beta) superfamily encompasses a large group of growth and differentiation factors playing important roles in regulating embryonic development and in maintaining tissue homeostasis in adult animals. Using degenerate polymerase chain reaction, we have identified a new murine TGF-beta family member, growth/differentiation factor-8 (GDF-8), which is expressed specifically in developing and adult skeletal muscle. During early stages of embryogenesis, GDF-8 expression is restricted to the myotome compartment of developing somites. At later stages and in adult animals, GDF-8 is expressed in many different muscles throughout the body. To determine the biological function of GDF-8, we disrupted the GDF-8 gene by gene targeting in mice. GDF-8 null animals are significantly larger than wild-type animals and show a large and widespread increase in skeletal muscle mass. Individual muscles of mutant animals weigh 2-3 times more than those of wild-type animals, and the increase in mass appears to result from a combination of muscle cell hyperplasia and hypertrophy. These results suggest that GDF-8 functions specifically as a negative regulator of skeletal muscle growth.
Article
Adipogenesis, or the development of fat cells from preadipocytes, has been one of the most intensely studied models of cellular differentiation. In part this has been because of the availability of in vitro models that faithfully recapitulate most of the critical aspects of fat cell formation in vivo. More recently, studies of adipogenesis have proceeded with the hope that manipulation of this process in humans might one day lead to a reduction in the burden of obesity and diabetes. This review explores some of the highlights of a large and burgeoning literature devoted to understanding adipogenesis at the molecular level. The hormonal and transcriptional control of adipogenesis is reviewed, as well as studies on a less well known type of fat cell, the brown adipocyte. Emphasis is placed, where possible, on in vivo studies with the hope that the results discussed may one day shed light on basic questions of cellular growth and differentiation in addition to possible benefits in human health.
Article
Myostatin, a new TGF-beta family member, is known as a muscle growth inhibitor, but its role in adipocyte development has not been studied. To test the role of Myostatin in 3T3-L1 preadipocyte differentiation, we treated cultured 3T3-L1 preadipocytes with Myostatin dissolved in 0.1% trifluoroacetic acid (TFA) during differentiation after they had become confluent. Myostatin treatment significantly decreased glycerol-3-phosphate dehydrogenase (GPDH) activity and oil Red-O staining compared to controls that did not receive Myostatin. Western blot analysis showed that the expression levels of CCAAT/enhancer binding protein alpha (C/EBP alpha) and peroxisome proliferator-activated receptor gamma (PPAR gamma) were significantly decreased by Myostatin treatment (P < 0.05). However, the expression of C/EBP beta was not significantly changed by the treatment (P > 0.05). From RT-PCR result, the relative level of leptin mRNA in Myostatin-treated cells was not significantly different (P > 0.1) from the level in cells without Myostatin treatment. Our data show that Myostatin, a secreted protein from muscle, inhibits preadipocyte differentiation in 3T3-L1 cells, which is mediated, in part, by altered regulation of C/EBP alpha and PPAR gamma.
Article
Growth differentiation factor-8 (GDF-8), or Myostatin, plays an important inhibitory role during muscle development. Since muscle and adipose tissue develop from the same mesenchymal stem cells, we hypothesized that Myostatin gene knockout may cause a switch between myogenesis and adipogenesis. Male and female wild type (WT) and Myostatin knockout (KO) mice were sacrificed at 4, 8, and 12 weeks of age. The gluteus muscle (GM) was larger in KO mice compared to WT mice at 8 (P < 0.01) and 12 (P < 0.001) weeks. At 12 weeks, KO mice had decreased fat depots (P < 0.01). Compared to 12-week-old WT mice, serum leptin concentration in KO mice was lower (P < 0.001) and leptin mRNA expression was decreased (P < 0.01) in inguinal adipose tissue. CCAAT/enhancer binding protein-alpha (C/EBPalpha) and peroxisome proliferator-activated receptor-gamma (PPARgamma) levels in adipose tissue were significantly lower in KO mice compared to WT mice. Thus, increased muscle development in Myostatin knockout mice is associated with reduced adipogenesis and consequently, decreased leptin secretion.