ArticlePDF Available

Gs signalling suppresses PPAR 2 generation and inhibits 3T3L1 adipogenesis

Authors:

Abstract and Figures

Since TSH receptor (TSHR) expression increases during adipogenesis and signals via cAMP/phospho-cAMP-response element binding protein (CREB), reported to be necessary and sufficient for adipogenesis, we hypothesised that TSHR activation would induce preadipocyte differentiation. Retroviral vectors introduced constitutively active TSHR (TSHR*) into 3T3L1 preadipocytes; despite increased cAMP (RIA) and phospho-CREB (western blot) there was no spontaneous adipogenesis (assessed morphologically, using oil red O and QPCR measurement of adipogenesis markers). We speculated that Gbetagamma signalling may be inhibitory but failed to induce adipogenesis using activated Gsalpha (gsp*). Inhibition of phosphodiesterases did not promote adipogenesis in TSHR* or gsp* populations. Furthermore, differentiation induced by adipogenic medium with pioglitazone was reduced in TSHR* and abolished in gsp* expressing 3T3L1 cells. TSHR* and gsp* did not inactivate PPARgamma (PPARG as listed in the HUGO database) by phosphorylation but expression of PPARgamma1 was reduced and PPARgamma2 undetectable in gsp*. FOXO1 phosphorylation (required to inactivate this repressor of adipogenesis) was lowest in gsp* despite the activation of AKT by phosphorylation. PROF is a mediator that facilitates FOXO1 phosphorylation by phospho-Akt. Its transcript levels remained constantly low in the gsp* population. In most measurements, the TSHR* cells were between the gsp* and control 3T3L1 preadipocytes. The enhanced down-regulation of PREF1 (adipogenesis inhibitor) permits retention of some adipogenic potential in the TSHR* population. We conclude that Gsalpha signalling impedes FOXO1 phosphorylation and thus inhibits PPARgamma transcription and the alternative promoter usage required to generate PPARgamma2, the fat-specific transcription factor necessary for adipogenesis.
Content may be subject to copyright.
Gsasignalling suppresses PPARg2 generation and inhibits 3T3L1
adipogenesis
Lei Zhang, Carol Paddon, Mark D Lewis, Fiona Grennan-Jones and Marian Ludgate
School of Medicine, Centre for Endocrine and Diabetes Sciences, Cardiff University, Heath Park, Cardiff CF14 4XN, UK
(Correspondence should be addressed to M Ludgate; Email: ludgate@cf.ac.uk)
Abstract
Since TSH receptor (TSHR) expression increases during
adipogenesis and signals via cAMP/phospho-cAMP-response
element binding protein (CREB), reported to be necessary
and sufficient for adipogenesis, we hypothesised that TSHR
activation would induce preadipocyte differentiation. Retro-
viral vectors introduced constitutively active TSHR
(TSHR*) into 3T3L1 preadipocytes; despite increased
cAMP (RIA) and phospho-CREB (western blot) there was
no spontaneous adipogenesis (assessed morphologically, using
oil red O and QPCR measurement of adipogenesis markers).
We speculated that Gbg signalling may be inhibitory but
failed to induce adipogenesis using activated Gsa(gsp*).
Inhibition of phosphodiesterases did not promote adipogen-
esis in TSHR* or gsp* populations. Furthermore, differen-
tiation induced by adipogenic medium with pioglitazone was
reduced in TSHR* and abolished in gsp* expressing 3T3L1
cells. TSHR* and gsp* did not inactivate PPARg(PPARG as
listed in the HUGO database) by phosphorylation but
expression of PPARg1 was reduced and PPARg2 undetect-
able in gsp*. FOXO1 phosphorylation (required to inactivate
this repressor of adipogenesis) was lowest in gsp* despite the
activation of AKT by phosphorylation. PROF is a mediator
that facilitates FOXO1 phosphorylation by phospho-Akt. Its
transcript levels remained constantly low in the gsp*
population. In most measurements, the TSHR* cells were
between the gsp* and control 3T3L1 preadipocytes. The
enhanced down-regulation of PREF1 (adipogenesis
inhibitor) permits retention of some adipogenic potential in
the TSHR* population. We conclude that Gsasignalling
impedes FOXO1 phosphorylation and thus inhibits PPARg
transcription and the alternative promoter usage required to
generate PPARg2, the fat-specific transcription factor
necessary for adipogenesis.
Journal of Endocrinology (2009) 202, 207–215
Introduction
Adipose tissue expands via two mechanisms, hypertrophy of
individual adipocytes and hyperplasia due to the proliferation
and differentiation of preadipocyte precursors (Drolet et al.
2008). In vitro protocols to induce preadipocyte differentiation
(adipogenesis) use agents to increase cAMP. Furthermore,
studies in the 3T3L1 murine preadipocytes, the cell line that
has been central in unravelling the complex mechanisms
driving adipogenesis, indicated that treatment with forskolin
or isobutylmethylxanthine (IBMX) alone suffices to trigger
the process (Boone et al. 1999) and it has been shown
subsequently that cAMP-response element binding protein
(CREB) activation is both necessary and sufficient to induce
adipogenesis (Reusch et al. 2000).
Many G protein-coupled receptors, including the thyro-
tropin receptor (TSHR), which is the main regulator of
thyroid function and growth, signal predominantly via the
cAMP/protein kinase A (PKA) pathway, ultimately leading
to phosphorylation of CREB (Vassart & Dumont 1992).
TSHR expression is also increased during adipogenesis
(Haraguchi et al. 1996). Does this suggest a role for TSHR
activation in adipocyte biology? The question is clinically
relevant since thyroid dysfunction is common, affecting up to
2% of the population, and the majority of these individuals
will have overactivation of the TSHR either by supraphy-
siological concentrations of TSH in hypothyroidism or from
thyroid stimulating antibodies in Graves’ hyperthyroidism
(Vanderpump et al. 1995).
On the basis of the known signalling pathways following
from TSHR activation, we hypothesise that it should trigger
adipogenesis spontaneously or at least enhance that induced
by known differentiating agents such as synthetic PPARg
(PPARG as listed in HUGO database) agonists. We have used
an in vitro model in which naturally occurring constitutively
active mutant forms of TSHR, found in e.g. toxic nodules
(Paschke & Ludgate 1997), are introduced using retroviral
vectors (Fuhrer et al. 2003). Our previous studies used the
model to investigate the effect of TSHR activation on human
orbital primary preadipocytes (of relevance to the eye disease
occurring in Graves’ patients). We demonstrated spontaneous
induction of early, but inhibition of the later stages of
adipogenesis, even in the presence of PPARgagonists (Zhang
et al. 2006). Orbital preadipocytes are a particular fat depot,
207
Journal of Endocrinology (2009) 202, 207–215 DOI: 10.1677/JOE-09-0099
0022–0795/09/0202–207 q2009 Society for Endocrinology Printed in Great Britain Online version via http://www.endocrinology-journals.org
This is an Open Access article distributed under the terms of the Society for Endocrinology’s Re-use Licence which permits unrestricted non-commercial use,
distribution, and reproduction in any medium, provided the original work is properly cited.
being derived from the neural crest, and this may explain the
differences between our results following activation of the
cAMP pathway and those reported using a murine
preadipocyte cell line. In the current study, we have applied
the model (activating mutant TSHR–L629F) to 3T3L1. To
our surprise, despite modestly increasing intracellular cAMP
and elevating phospho-CREB levels, TSHR activation did
not induce adipogenesis and this was not changed by adding a
phosphodiesterase inhibitor. Furthermore, adipogenesis
induced by PPARgagonists was significantly reduced in the
mutant TSHR expressing cells compared with the non-
modified population. Since TSHR activation liberates two
functional moieties of the Gs protein, a(which acts via PKA)
and bg (signals via PI3K), we suggested that the inhibitory
effects were due to the latter. Thus, we have introduced
constitutively active Gsa(gsp*), Q227L (Ludgate et al. 1999),
which yields only the asubunit. Again there was no
spontaneous adipogenesis and in vitro induced differentiation
was completely abolished. Further investigations revealed a
reduction in PPARg1 and the complete absence of PPARg2
proteins, the fat-specific isoform, in gsp* expressing 3T3L1.
Materials and Methods
Reagent source; cell culture and adipogenesis protocols
All chemicals were obtained from Sigma–Aldrich and tissue
culture media and serum from BioWhittaker-Lonza
(Verviers, Belgium) unless otherwise stated. The 3T3L1 cell
line was purchased from the ATCC (Atlanta, GA, USA).
3T3L1 murine preadipocytes were routinely cultured in
DMEM/F12 10% FCS (complete medium, CM). Adipogen-
esis was induced in confluent cells by replacing with
differentiation medium (DM) containing 5% FCS, biotin
(33 mM), panthothenate (17 mM), tri-iodothyronine (1 nM),
dexa-methasone (100 nM), thiazolidinedione (1 mM) and
insulin (500 nM), for 10–12 days, as previously described
(Zhang et al. 2006).
Activating mutant human TSHR, L629F (TSHR*) and rat
Gsa, Q227L (gsp*), were introduced using retroviral vectors,
previously produced in our laboratory (Ivan et al. 1997,
Fuhrer et al. 2003). All experiments were performed on the
mixed pools of cells that resulted from geneticin selection.
Furthermore, at least two independent mixed pools were
generated for each cell population, with results being
comparable. Initial characterisation of the G418 resistant
pools of 3T3L1 employed RT-PCR and direct sequencing to
demonstrate the expression of human TSHR and rat Gsain
the appropriate populations as previously described (Ivan et al.
1997,Fuhrer et al. 2003). In many of our experiments we also
included 3T3L1 populations transduced with the wild-type
(WT) and a second mutant form of TSHR, M453T. Results
obtained (basal levels of cAMP/pCREB, transcript measures
for PPARgand GPDH, etc.) indicated that there was no
significant difference in the behaviour of WT compared
with non-modified or M453T compared with L629F. In the
interests of brevity, the results are reported for the non-
modified, L629F and gsp* expressing populations.
Effect on cAMP levels
The non-modified 3T3L1 and populations expressing gsp*or
TSHR* were plated at 5!10
4
/well in 12-well plates. The
following day they were incubated for 4 h in medium
containing 10
K4
M IBMX with no further addition (basal
conditions), or plus 10 mU/ml bovine TSH. Forskolin (in
addition to IBMX) at 10
K5
M was included as a positive
control in all populations. cAMP was extracted in 0.1 M HCl
and measured using an in-house RIA capable of detecting
femtogram quantities of the second messenger, as previously
described (Fuhrer et al. 2003). Coulter counting of adjacent
wells provided an accurate cell number for results to be
expressed as picomoles cAMP/10
4
cells.
Western blotting
The three different 3T3L1 populations were propagated in
duplicate in 6-well plates in CM or DM. Proteins were
extracted, at various time points, in Laemmli buffer
containing 1 mM sodium orthovanadate and 1 mM phenyl-
methylsulphonyl fluoride. To prepare nuclear and cytosolic
fractions, cells were harvested in ice-cold PBS, centrifuged
and resuspended in HEPES buffer containing protease
inhibitors. The supernatant produced by centrifugation at
12 000 gprovided the cytosolic and high-salt extraction of the
pellet, the nuclear fractions respectively.
Samples (containing 20 mg protein) were separated by 10%
SDS-PAGE and then the gel electroblotted onto PVDF
membrane as previously described (Al-Khafaji et al. 2005).
To investigate various signalling pathways, the blots were
probed with the following rabbit antibodies (all from Cell
Signalling Technology unless otherwise stated): anti
phospho-CREB (Ser 133, 1:2000 overnight; 4 8C); anti
total-CREB (1:1000, room temperature, 1 h; Santa Cruz
Biotechnology, Santa Cruz, CA, USA); anti phospho-Akt
(Thr 308, 1:1000, 4 8C, overnight); anti total-Akt (1:1000,
room temperature, 1 h, Santa Cruz); and anti phospho-SYK
(Tyr 352, 1:5000, overnight, 4 8C, BD Biosciences, San Jose,
CA, USA). To investigate transcription factors, rabbit
antibodies to phospho-PPARg(Ser 82, 1:1000 overnight,
48C, Upstate), anti total-PPARg(1:1000, overnight, 4 8C),
anti total-FOXO1 (1:1000, overnight, 4 8C) and anti
phospho-FOXO1 (Ser 256, 1:1000, overnight, 4 8C) were
employed. In all cases proteins were detected using either an
anti-mouse IgG-HRP conjugate or an anti-rabbit IgG-HRP
conjugate (1:5000, room temperature for 1 h, GE Health-
care, Amersham, UK) and visualised by enhanced chemi-
luminescence (ECL Plus, GE Healthcare).
Films were analysed using the Alpha Imager 1200
digital imaging system (Alpha Innotech Corp., San Leandro,
CA, USA). The blots were initially probed with the
L ZHANG and others .Gsasignalling, PPARg2 and adipogenesis208
Journal of Endocrinology (2009) 202, 207–215 www.endocrinology-journals.org
phospho-specific antibodies; they were then stripped and
reprobed with antibodies that recognise total proteins.
Effect on spontaneous and PPARginduced adipogenesis
The various cell populations (in 12-well plates) were
examined in CM and DM. Microscopic examination
provided a means of determining whether morphological
changes, for e.g. rounding-up of cells and/or acquisition of
lipid-filled droplets (oil red O staining), had occurred.
Effect on mitotic clonal expansion
To investigate whether the expression of the various mutants
had any effect on the mitotic clonal expansion phase, reported
by some authors to be required for differentiation in 3T3L1
preadipocytes (Tang et al. 2003), the cells were counted 1, 2, 3
and 10 days after addition of DM, using a Coulter particle
counter; results are expressed (meanGS.E.M.) fold increase in
cell number.
QRT-PCR measurement of transcript copy number
The various cell populations were plated in 6-well plates in
CM or DM. Ten days later, RNA was extracted, reverse
transcribed and transcript copy numbers for PPARg,GPDH,
PREF1,ID2,PROF, and acidic ribosomal phosphoprotein
(ARP) were measured using Sybr green and a Stratagene
MX3000 light cycler. Primers used are listed in Ta b l e 1 .
Standard curves (the PCR amplicon subcloned into
pGEM-T at 10
6
to 10
2
copies) were included for each gene
and results are expressed relative to the housekeeping gene
ARP. This gene was selected on the basis of its high expression
level (C
t
value 16G0.29 in non-modified 3T3L1 in CM),
which was not modified in the L629F and gsp* populations
(16G0.31 and 16G0.32 respectively in CM) or during
adipocyte differentiation (16G0.28 in non-modified cells on
day 9 in DM; the C
t
values are the meanGS.E.M. obtained
from four separate experiments).
Statistical analysis
Means were compared using Student’s t-test for parametric
data and non-parametric data were analysed using the
Wilcoxon signed ranks test.
Results
TSHR* and gsp* increase unstimulated cAMP levels
In non-modified 3T3L1, the unstimulated level of cAMP was
397G18 pmol/10
4
cells (results are the meanGS.E.M. of three
experiments all performed in at least triplicate). Unstimulated
cAMP levels were significantly increased in cells expressing
TSHR* or gsp* reaching 129G3.9% (P!0.02) and
140G5.9% (P!0.01) respectively of the non-modified
population. TSH (1 mU/ml) elicited a cAMP response in
cells expressing mutant TSHR (650G218% of unstimulated)
but not in the non-modified or gsp* 3T3L1 cells, in keeping
with the very low level of endogenous TSHR in these cells
prior to differentiation (Haraguchi et al. 1996). There was no
significant difference in the response to forskolin, which
induced a robust O20-fold increase in cAMP in all the three
populations.
Phosphorylated CREB levels are increased in TSHR* and gsp*
expressing cells and are modulated during adipogenesis
Western blots to investigate activation of CREB by
phosphorylation are usually performed following acute
stimulation of the cells followed by analysis in the minutes
after exposure to the stimulant. By contrast, we have
examined the basal levels of phospho-CREB in the three
populations of 3T3L1 to assess the effects of chronic activation
via the various mutants. We have used a value of 1 for the
phosphorylated:total CREB ratio of the non-modified
preadipocytes. Based on at least seven experiments using
two independently generated populations of TSHR* or gsp*
expressing cells, there was a modest, but significant increase in
the ratio in L629F (1.5G0.58 P!0.02) and gsp*(1
.19G0.22
P!0.02), reflecting chronic activation of this pathway.
A representative western blot is shown in Fig. 1.
To investigate whether CREB phosphorylation is affected
by the adipogenic process, we have measured the phosphor-
ylated:total CREB ratios in samples obtained from the
different 3T3L1 populations at various time points following
addition of DM. In the non-modified cells there is a slight
reduction in the ratio in mid-differentiation, but a significant
increase in the later stages (P!0.05), to achieve levels
150–300% of those on day 0 (prior to addition of DM). Apart
from the higher basal levels of phospho-CREB, a similar
Table 1 Primer sequences (and amplicon sizes) employed for QPCR measurements
Forward primer Reverse primer
Gene
PPARg220 bp TTTTCAAGGGTGCCAGTTTC AATCCTTGGCCCTCTGAGAT
GPDH 124 bp ATGCTCGCCACAGAATCCACAC AACCGGCAGCCCTTGACTTG
PREF1 148 bp CGTGATCAATGGTTCTCCCT AGGGGTACAGCTGTTGGTTG
ARP 72 bp GAGGAATCAGATGAGGATATGGGA AAGCAGGCTGACTTGGTTGC
PROF 72 bp GATCACTCTGTCATCATGTGG CTTACTGTCCCACATTTGCTTG
ID2 141 bp GGACCACAGCTTGGGCAT CGTTCATGTTGTAGAGCAGACTCAT
Gsasignalling, PPARg2 and adipogenesis .L ZHANG and others 209
www.endocrinology-journals.org Journal of Endocrinology (2009) 202, 207–215
pattern of expression was apparent in the TSHR* and gsp*
expressing cells through adipogenesis (data not shown).
The increased phospho-CREB levels do not produce spontaneous
adipogenesis
In CM, even when the cells had been confluent for up to
10 days, there were no morphological changes consistent with
adipogenesis in cells expressing gsp*orTSHR*when
compared with the control 3T3L1. Occasional cells containing
small lipid vacuoles were observed in all populations.
Furthermore, we did not see a consistent change in
transcript levels of PREF1,PPARgor GPDH in cells
expressing TSHR* or gsp* when compared with the non-
modified population (data not shown).
Since chronic stimulation of the cAMP pathway might
induce up-regulation of e.g. phosphodiesterases, we included
a phosphodiesterase inhibitor (IBMX) in the culture medium.
Cells were allowed to reach confluence in CM and 0.5mM
IBMX was added for various time periods, but there was still
no morphological evidence for adipogenesis and the
transcript levels for markers of differentiation were not
significantly changed (data not shown).
TSHR* and gsp* inhibit induced adipogenesis by affecting
PPARgisoforms
PPARgagonist induced differentiation of the non-modified
cells produced the expected change in morphology and
similar signs of adipogenesis were also present in the L629F
expressing cells, but at a reduced level compared with the
control; by contrast, the gsp* population were completely
devoid of differentiating cells, as illustrated in Fig. 2, by the
absence of oil red O stained cells.
We compared transcr ipts for markers of adipogenesis in
non-modified, L629F and gsp* expressing cells; statistical
analyses of the results reported as fold changes are shown in
Table 2.Figure 3 is a representative experiment illustrating
that PPARgagonist induced differentiation of non-modified
3T3L1 resulted in sustained and significant increases in
PPARgand GPDH (P!0.05 and 0.02 respectively).
The L629F and gsp* populations displayed an attenuated
increase in PPARg(although not significantly different from
non-modified) and in GPDH (both significantly less than
non-modified) being more severe in the gsp* expressing cells
in keeping with their morphological appearance. Further-
more, expression of PREF1, an EGF-like transmembrane
protein that inhibits adipogenesis (Smas & Sul 1993), is
significantly down-regulated (P!0.04), by the differentiation
Figure 1 Representative (one out of seven performed) western blot of
the three 3T3L1 populations in complete medium (basal conditions)
showing phosphorylated CREB (upper panel) and total CREB (lower
panel) both having an apparent molecular weight of 43 kDa.
Figure 2 Oil red O staining in 3T3L1 cells following nine
days in differentiation medium containing pioglitazone.
A, non-modified; B, L629F; and C, gsp* expressing populations.
Magnification !200.
L ZHANG and others .Gsasignalling, PPARg2 and adipogenesis210
Journal of Endocrinology (2009) 202, 207–215 www.endocrinology-journals.org
protocol, to 54G0.12% of the day 0 value, in the non-
modified cells. The behaviour of PREF1 in the other
populations diverges. PREF1 expression in L629F cells after
exposure to DM was reduced to 30G0.06% of day 0 levels
(P!0.005), and this differentiation induced PREF1 down-
regulation is significantly greater than in the non-modified
and gsp* populations (P!0.04 and 0.02 respectively). By
contrast, the reduction to 89G0.07% of pre-treatment values
in the gsp* population is not significantly different from day 0
PREF1 transcripts in CM in these cells.
Coulter counting of the cells in the first 3 days after
addition of DM (nZ3) revealed that the non-modified
3T3L1 had a 2.4G0.9-fold increase in cell number on day 3
compared with day 0, indicating mitotic clonal expansion
(MCE). Both L629F and gsp* populations displayed
significantly increased proliferation compared with the non-
modified, 4.9G0.5(P!0.005) and 3.3G0.7(P!0.01) fold
respectively, suggesting no influence of either mutation on
MCE. In the following seven days, there was no further
increase in either the non-modified or L629F cell number,
but the gsp* population continued to proliferate, in keeping
with the absence of differentiation.
The transcriptional activity of PPARgis reduced when it is
phosphorylated (Hu et al. 1996). We were unable to detect
any phosphorylated PPARgeither in total cell lysates or
nuclear extracts. However, as shown in Fig. 4, the expression
of PPARg1 is greatly reduced (at all time points) in L629F
and gsp* and PPARg2 is completely absent from the latter, in
contrast to the non-modified 3T3L1.
Furthermore, the expression of PPARg1 and PPARg2
proteins in non-modified cells is increased in the first 24 h
following induction. PPARg1 expression continues to
increase throughout adipogenesis, but PPARg2 expression
is at the limit of detection during the MCE stage, but then
resumes in the terminal stages of differentiation.
Reduced FOXO1 phosphorylation may explain the lack of
PPARg2
The transcription factor FOXO1 represses the promoters for
PPARg1 and PPARg2(Armoni et al. 2006); it is inactivated
by phosphorylation when it translocates from the nucleus
to the cytoplasm (Nakae et al. 2003). As can be seen in
Table 2 Fold changes in transcript levels of adipogenesis markers in
the three populations of 3T3L1 cells following exposure to
differentiation medium. The three populations were plated in
12-well plates; once confluent, the cells were cultured for 9 days in
differentiation medium containing pioglitazone. mRNA was
extracted on days 0 and 9 and QPCR measurement of the
adipogenesis markers was performed. Results (meanGS.E.M. of the
three experiments were all performed in at least duplicate) are the
fold changes in transcripts for each gene (relative to the ARP
housekeeper) comparing day 9 and day 0
Non-modified
cells L629F population gsp* population
Marker
PPARg6.5G2.52
.7G0.72
.2G0.5
GPDH 159G43 19.5G8* 2.2G0.43
PREF1 0.54G0.12 0.3G0.06
0.89G0.07
§
*P!0.03 compared with non-modified;
P!0.02 compared with non-
modified;
P!0.04 compared with non-modified;
§
P!0.02 compared
with L629F.
Figure 3 QPCR measurement of adipogenesis markers – A, PREF1;
B, PPARg; and C, GPDH on day 0 (cells in complete medium) and
day 9 (cells in differentiation medium containing pioglitazone) in
non-modified, L629F and gsp* 3T3L1 cells. Results are the
meansGS.E.M. of triplicates, expressed as absolute transcript copy
numbers (transcripts) of the gene of interest per 1000 copies (100 for
PREF-1) of acidic ribosomal phosphoprotein (ARP). Representative
experiment, one out of three performed.
Gsasignalling, PPARg2 and adipogenesis .L ZHANG and others 211
www.endocrinology-journals.org Journal of Endocrinology (2009) 202, 207–215
Fig. 5, in the non-modified cells total FOXO1 protein
expression increases in the first 24 h after addition of DM in
the nucleus, where it is highly phosphorylated. It is essentially
absent from the cytoplasm during the MCE phase. In the gsp*
population, total FOXO1 protein expression is higher in basal
conditions, its expression is increased in DM and we observe
its translocation from the cytoplasm to the nucleus between
days 1 and 3. The main difference is the considerably
decreased level of phosphorylation from day 3 onwards
compared with the non-modified 3T3L1. The TSHR* cells’
behaviour is midway between that of the other two
populations.
What accounts for the behaviour of FOXO1?
Phosphorylation-dependent inactivation of FOXO1 depends
chiefly on the PI3K and phospho-Akt pathway (Sakaue et al.
1998). We investigated and observed an early increase in the
proportion of phospho-Akt in all populations of 3T3L1 cells
after addition of DM (Fig. 6), indicating the need for an
alternative explanation for the reduced FOXO1 phosphoryl-
ation in gsp*.
Two novel modulators of adipogenesis have recently been
described. PROF is a mediator between AKT and FOXO1
whose expression is transiently up-regulated during adipo-
genesis (Fritzius & Moelling 2008). ID2 is a small molecule,
whose transcription also increases during adipogenesis and
interacts with as yet to be identified factors to stimulate
PPARgexpression (Park et al. 2008). We investigated their
expression and found a transient increase in ID2 TCN in all
three populations in the first 4–8 h following addition of
DM, although transcript levels of ID2 were lowest in gsp*
cells at all time points (data not shown). By contrast, as shown
in Fig. 7,PROF transcription was increased in the non-
modified and, to a lesser extent, in the L629F populations, but
remained at a constant low level in the gsp* cells. Thus,
despite the abundance of phospho-Akt in the gsp* population,
the paucity of PROF prevents inactivation of FOXO1 by
phosphorylation, PPARg2 is not expressed and adipogenesis
does not occur.
Discussion
Our initial aim was to investigate the effects of TSHR
activation on 3T3L1 preadipocytes, following our previous
studies demonstrating increased adipogenesis in ex vivo fat
samples from Graves’ patients, in whom thyroid stimulating
antibodies (as opposed to gain-of-function mutation) produce
chronic activation of the TSHR (Starkey et al. 2003). From
reports of a pro-adipogenic effect of agents capable of
elevating cAMP, we hypothesised that spontaneous adipogen-
esis would arise in preadipocytes expressing naturally
occurring (e.g. in toxic thyroid adenoma) gain-of-function
Figure 4 Western blot analysis of PPARgprotein expression on day 0 (cells in complete medium) and at various time points
following addition of differentiation medium containing pioglitazone in non-modified, L629F and gsp* 3T3L1 cells.
Representative experiment, one out of three performed.
Figure 5 Western blot analysis of FOXO1 protein expression on day 0 (cells in complete medium) and at various time points
following addition of differentiation medium containing pioglitazone in non-modified, L629F and gsp* 3T3L1 cells. A, Nuclear
extracts probed with anti-phospho-FOXO1 (upper panel) and anti-total FOXO1 (lower panel). B, cytosolic extracts (panels as in
A). Representative experiment, one out of three performed.
L ZHANG and others .Gsasignalling, PPARg2 and adipogenesis212
Journal of Endocrinology (2009) 202, 207–215 www.endocrinology-journals.org
TSHR mutations (which like thyroid stimulating antibodies,
activate adenylate cyclase/PKA), but this was not the case.
It should be noted that the increase in cAMP in preadipocytes
using retroviral vectors to express the mutant receptor is
considerably lower than that obtained following overexpres-
sion by transfection of COS cells. The model more closely
resembles the in vivo situation, in which modest elevation of
cAMP is sufficient to generate a thyroid toxic nodule, and
concurs with our previous findings in thyroid cells (Ludgate
et al. 1999,Fuhrer et al. 2003). The constitutively active
mutant TSHR responded further when stimulated with TSH,
in common with other such mutants, although its TSH
responsiveness was less than the WT TSHR (data not shown)
as reported by others http://innere.uniklinikum-leipzig.de/
tsh/frame.html.
We have then used a gain-of-function mutation in a second
component of the PKA pathway, Gsasubunit, to achieve
activation of CREB. Neither the receptor, nor its down-
stream G protein, induced spontaneous adipogenesis (even in
the presence of IBMX) as would be expected from previous
reports indicating that CREB activation is ‘necessary and
sufficient for adipogenesis in 3T3L1’ (Reusch et al. 2000).
Experiments were then conducted to investigate the effects
of TSHR* and gsp* on in vitro induced adipogenesis.
The reduced differentiation in 3T3L1 expressing TSHR*
mirror the results we obtained in human primary orbital
preadipocytes (Zhang et al. 2006). Our demonstration of an
inhibitory role for Gsain adipogenesis confirms and extends
the findings of Wa n g et al. (1992) who demonstrated
enhanced differentiation in 3T3L1 cells treated with
oligomers antisense to the protein. Subsequently, they
(Wang & Malbon 1999) reported that Gsarepression of
adipogenesis is mediated by the tyrosine kinase Syk, but how
this may impact adipogenesis is not clear and we are not aware
of further studies to address this question. Our investigations
of the pathway confirmed their findings, i.e. an early and
transient increase in phospho-SYK in the first day following
induced adipogenesis, but only in the non-modified
population. The L629F and gsp* expressing cells showed a
decrease in phospho-SYK levels (data not shown).
Our experiments illustrate that the failure to differentiate in
gsp* is due to reduced expression of PPARg1 and interference
in the alternate promoter usage required to generate
PPARg2, the isoform specific to adipose tissue (Tontonoz
et al. 1995).
There is some controversy concerning how PPARg2
expression changes following addition of DM. Some authors
report that it is induced, others that it is up-regulated. In our
differentiation protocol; PPARg2 protein expression
increases in the first 24 h, but is then repressed until day 5
post-induction. Its reappearance seems to coincide with the
start of the terminal differentiation phase and was a consistent
finding in all experiments, but contrasts with the results
obtained by others reporting that PPARg2 expression steadily
increases through all stages of differentiation (Saladin et al.
1999).
FOXO1 represses transcription from both the PPARg1
and PPARg2 promoters (Armoni et al. 2006) and so we
investigated its participation in our different 3T3L1 popu-
lations. We observed a rapid increase in FOXO1 protein
expression and its subsequent translocation to the nucleus
during the first 3 days post-induction in non-modified
3T3L1. At later time points, the nuclear FOXO1 sustained a
high degree of phosphorylation. The antibody we have used
detects phosphorylation of FOXO1 at serine 256, located in
the forkhead domain and indicative of inactivation of the
transcription factor; there are two additional phosphorylation
sites. Nakae et al. (2003) transduced cells with a tagged
Figure 6 (A) Western blot analysis of phospho-Akt (upper panel) and actin (lower panel) in nuclear extracts on day 0 (cells in
complete medium) and at various time points following addition of differentiation medium containing pioglitazone in
non-modified, L629F and gsp* 3T3L1 cells.
Figure 7 QPCR measurement of ProF expression on day 0 (cells in
complete medium) and at various time points following the addition
of differentiation medium containing pioglitazone in non-modified,
L629F and gsp* 3T3L1 cells. Results are the meansGS.D.of
triplicates, expressed as absolute transcript copy numbers
(transcripts) per 1000 copies of acidic ribosomal phosphoprotein
(ARP). Representative experiment one out of two performed.
Gsasignalling, PPARg2 and adipogenesis .L ZHANG and others 213
www.endocrinology-journals.org Journal of Endocrinology (2009) 202, 207–215
FOXO1 and were able to demonstrate a cytoplasmic location
for phosphorylated FOXO1. In our experiments investigating
endogenous FOXO1, it is predominantly cytoplasmic, then
nuclear and then present in both fractions, in agreement with
this work. Any phosphorylated endogenous FOXO1 in our
nuclear extracts is probably partially phosphorylated (on
serine 256, the preferred target for the kinase) and thus
inactive. Our experiment also revealed that phosphorylation
of FOXO1 is impeded in the gsp* population from day 3 post-
induction onwards and so provides some explanation for their
lack of adipogenesis.
At no time point is signalling via phospho-Akt compro-
mised and yet FOXO1 phosphorylation had remained low in
the gsp* cells. Investigation of two recently described
modulators of adipogenesis, ID2 and PROF suggested they
may have a role in explaining this anomaly. PROF, also known
as WDFY2, is a member of the WD-repeat propeller-FYVE
protein family, which is a binding partner for phospho-Akt
and facilitates FOXO1 phosphorylation (Fritzius & Moelling
2008). In silico analysis of its proximal promoter reveals the
presence of a potentially functional half-site cAMP response
element; however, it lacks a ‘tata’ box, generally thought to be
required for robust transcriptional induction by cAMP
(Zhang et al. 2005). Our investigations demonstrate that its
expression level, although higher in basal conditions,
remained fairly constant at all time points in the gsp* cells
following addition of DM, in contrast to the marked increase
in the non-modified 3T3L1.
We also considered the possibility that the excess Gsain the
gsp* expressing cells might sequester free Gbg subunits
generated during adipogenesis and thus impede their down-
stream effects. However, the mutation in arenders it
permanently in the GTP-bound form and thus less likely to
bind bg (Ford et al. 1998).
In the majority of our experiments, the behaviour of the
TSHR* population is mid-way between that of the gsp*
expressing and parent cell lines, indicating that the inhibitory
effects of Gsaare abrogated, presumably by the b/gsubunits,
which in the thyroid have been reported to signal by PI3K
(Zaballos et al.2008). PREF1 inhibits adipogenesis by
preventing induction of PPARg2(Kim et al. 2007). The
significantly enhanced down-regulation of PREF1 in the
TSHR* expressing cells may contr ibute to the rescue
mechanism. The drawing depicted in Fig. 8 summarises the
mechanism we propose to explain the difference in
adipogenic potential of the TSHR* and gsp* expressing
3T3L1 populations.
In conclusion, our results identify an important role for
Gsasignalling in inhibiting the alternative promoter usage
necessary to produce PPARg2transcripts.
Declaration of interest
There is no conflict of interest that could be perceived as prejudicing the
impartiality of the research reported.
Funding
Supported by The Wellcome Trust (grant number WT076003MA).
Acknowledgements
We would like to thank Prof. Maurice Scanlon for his continued support.
References
Al-Khafaji F, Wiltshire M, Fuhrer D, Mazziotti G, Lewis MD, Smith PJ
& Ludgate M 2005 Biological activity of activating TSHR mutants:
modulation by iodide. Journal of Molecular Endocrinology 34 209–220.
Armoni M, Harel C, Karni S, Chen H, Bar-Yoseph F, Ver MR,
Quon MJ & Karnieli E 2006 FOXO1 represses peroxisome
proliferator activator receptor gamma 1 and gamma 2 gene promoters
in primary adipocytes – a novel paradigm to increase insulin
sensitivity. JournalofBiologicalChemistry28 19881–19891.
Boone C, Gregoire F & Remacle C 1999 Var ious stimulators of the cyclic
AMP pathway fail to promote adipose conversion of porcine preadipocytes
in primary culture. Differentiation 64 255–262.
Drolet R, Richard C, Sniderman AD, Mailloux J, Fortier M, Huot C,
Rheaume C & Tchernov A 2008 Hypertrophy and hyperplasia of
abdominal adipose tissues in women. International Journal of Obesity
32 283–291.
Ford CE, Skiba NP, Bae H, Daaka Y, Reuveny E, Shekter LR, Rosal R,
Weng G, Yang C-S, Iyengar R et al. 1998 Molecular basis for interactions of
G protein bg subunits with effectors. Science 280 1271–1274.
Fritzius T & Moelling K 2008 Akt- and Foxo1-interacting WD-repeat-FYVE
protein promotes adipogenesis. EMBO Journal 27 1399–1410.
Fuhrer D, Lewis MD, Al-Khafaji F, Starkey K, Paschke R, Wynford-Thomas
D, Eggo M & Ludgate M 2003 Biological activity of activating TSHR
mutants depends on the cellular context. Endocrinology 144 4018–4030.
Haraguchi K, Shimura H, Lin L, Saito T, Endo T & Onaya T 1996 Functional
expression of thyrotropin receptor in differentiated 3T3-L1 cells: a possible
model cell line of extrathyroidal expression of thyrotropin receptor.
Biochemical and Biophysical Research Communications 223 193–198.
Hu E, Kim JB, Sarraf P & Spiegelman BM 1996 Inhibition of adipogenesis
through MAP kinase-mediated phosphorylation of PPAR. Science 274
2100–2103.
Ivan M, Ludgate M, Gire V, Bond J & Wynford-Thomas D 1997 An
amphotropic retroviral vector expressing a mutant Gsp oncogene: effects on
human thyroid cells in vitro.Journal of Clinical Endocrinology and Metabolism
82 2702–2709.
Figure 8 A drawing summarising the effects of signalling via Gsa
and Gbg subunits in regulating adipogenesis via expression of
PPARg2.
L ZHANG and others .Gsasignalling, PPARg2 and adipogenesis214
Journal of Endocrinology (2009) 202, 207–215 www.endocrinology-journals.org
Kim KA, Kim JH, Wang Y & Sul HS 2007 Pref-1 (preadipocyte factor 1)
activates the MEK/extracellular signal-regulated kinase pathway to inhibit
adipocyte differentiation. Molecular and Cellular Biology 27 2294–2308.
Ludgate M, Gire V, Crisp M, Ajjan R, Weetman A, Ivan M & Wynford-
Thomas D 1999 Contrasting effects of Gsp and activating mutations of the
thyrotropin receptor on the function and proliferation of thyroid follicular
cells. Oncogene 18 4798–4807.
Nakae J, Kitamura T, Kitamura Y, Biggs WH III, Arden KC & Accili D 2003
The forkhead transcription factor Foxo1 regulates adipocyte differentiation.
Developmental Cell 4119–129.
Park KW, Waki H, Villanueva CJ, Monticelli LA, Hong C, Kang S,
MacDougald OA, Goldrath AW & Tontonoz P 2008 Inhibitor of DNA
binding 2 is a small molecule-inducible modulator of peroxisome
proliferator-activated receptor-gexpression and adipocyte differentiation.
Molecular Endocrinology 22 2038–2048.
Paschke R & Ludgate M 1997 The thyrotropin receptor and thyroid disease.
New England Journal of Medicine 337 1675–1681.
Reusch JEB, Colton LA & Klemm DJ 2000 CREB activation induces
adipogenesis in 3T3-L1 cells. Molecular and Cellular Biology 20 1008–1020.
Sakaue H, Ogawa W, Matsumoto M, Kuroda S, Takata M, Sug imoto T,
Spiegelman BM & Kasuga M 1998 Posttranscriptional control of adipocyte
differentiation through activation of phosphoinositide 3-kinase. Journal of
Biological Chemistry 273 28945–28952.
Saladin R, Fajas L, Dana S, Halvorsen YD, Auwerx J & Briggs M 1999
Differential regulation of peroxisome proliferator activator receptor gamma
1 (PPAR gamma 1) and PPAR gamma 2 messenger RNA expression in the
early stages of adipogenesis. Cell Growth & Differentiation 10 43–48.
Smas CM & Sul HS 1993 Pref-1, a protein containing EGF-like repeats,
inhibits adipocyte differentiation. Cell 73 725–734.
Starkey KJ, Janezic A, Jones G, Jordan N, Baker G & Ludgate M 2003 Adipose
thyrotropin receptor expression is elevated in Graves’ and thyroid eye
diseases ex vivo and indicates adipogenesis in progress in vivo.Journal of
Molecular Endocrinology 30 369–380.
Tang QQ, Otto TC & Lane MD 2003 Mitotic clonal expansion:
a synchronous process required for adipogenesis. PNAS 100 44–49.
Tontonoz P, Hu E, Devine J, Beale EG & Spiegelman BM 1995 PPAR2
regulates adipose expression of the phosphoenolpyruvate carboxykinase
gene. Molecular and Cellular Biology 15 351–357.
Vanderpump MPJ, Tunbridge WMG, French JM, Appleton D, Bates D,
Clark F, Gr imley J Evans, Hasan M, Rodgers H, Tunbridge F et al. 1995
The incidence of thyroid disorders in the community: a twenty-year
follow-up of the Whickham Survey. Clinical Endocrinology 43 55–68.
Vassart G & Dumont JE 1992 The thyrotropin receptor and the regulation of
thyrocyte function and growth. Endocrine Reviews 13 596–611.
Wang HY & Malbon CC 1999 Gs alpha inhibition of adipogenesis is via SYK.
Journal of Biological Chemistry 274 32159–32166.
Wang HY, Watkins DC & Malbon CC 1992 Antisense oligodeoxynucleotides
to G
s
proteina-subunit sequence accelerate differentiation of fibroblasts to
adipocytes. Nature 358 334–337.
Zaballos MA, Garcia B & Santisteban P 2008 Gbg dimers released in response
to TSH activate phosphoinositide 3 kinase and regulate gene expression in
thyroid cells. Molecular Endocrinology 16 342–352.
Zhang X, Odom DT, Koo S-H, Conkright MD, Canettieri G, Best J,
Chen H, Jenner R, Herbolsheimer E, Jacobsen E et al. 2005
Genome-wide analysis of cAMP-response element binding protein
occupancy, phosphorylation, and target gene activation in human tissues.
PNAS 102 4459–4464.
Zhang L, Baker G, Janus D, Paddon C, Fuhrer D & Ludgate M 2006
Biological effects of thyrotropin receptor activation on h uman orbital
preadipocytes. Investigative Ophthalmology & Visual Science 47
5197–5203.
Received in final form 29 April 2009
Accepted 21 May 2009
Made available online as an Accepted Preprint
21 May 2009
Gsasignalling, PPARg2 and adipogenesis .L ZHANG and others 215
www.endocrinology-journals.org Journal of Endocrinology (2009) 202, 207–215
... We and others have demonstrated that activation of the thyrotropin receptor (TSHR) in orbital preadipocyte-fibroblasts (OF) leads to increase in adipogenesis and hyaluronan production [12,13]. During adipogenesis, TSHR expression has been shown to increase [5] but little is known about the effects of TSHR activation at various differentiation stages. ...
... Cell lysates from OF at various time points before and during adipogenesis were obtained by addition of lysis buffer as previously described [13]. The culture supernatants from the same time points were also collected and concentrated using spin columns (Merck Millipore, Watford, Hertfordshire, UK) to produce an 80-fold concentration. ...
... The culture supernatants from the same time points were also collected and concentrated using spin columns (Merck Millipore, Watford, Hertfordshire, UK) to produce an 80-fold concentration. Lysates and concentrated supernatants were separated using SDS-PAGE as previously described [13]. Briefly, proteins were extracted, at various time points, in Laemmli buffer containing 1 mM phenylmethylsulphonyl fluoride. ...
Article
Full-text available
Thyroid stimulating antibodies (TSAB) cause Graves’ disease and contribute to Graves’ Orbitopathy (GO) pathogenesis. We hypothesise that the presence of TSH binding proteins (truncated TSHR variants (TSHRv)) and/or nonclassical ligands such as thyrostimulin (α2β5) might provide a mechanism to protect against or exacerbate GO. We analysed primary human orbital preadipocyte-fibroblasts (OF) from GO patients and people free of GO (non-GO). Transcript (QPCR) and protein (western blot) expression levels of TSHRv were measured through an adipogenesis differentiation process. Cyclic-AMP production by TSHR activation was studied using luciferase-reporter and RIA assays. After differentiation, TSHRv levels in OF from GO were significantly higher than non-GO (p = 0.039), and confirmed in ex vivo analysis of orbital adipose samples. TSHRv western blot revealed a positive signal at 46 kDa in cell lysates and culture media (CM) from non-GO and GO-OF. Cyclic-AMP decreased from basal levels when OF were stimulated with TSH or Monoclonal TSAB (M22) before differentiation protocol, but increased in differentiated cells, and was inversely correlated with the TSHRv:TSHR ratio (Spearman correlation: TSH r = −0.55, p = 0.23, M22 r = 0.87, p = 0.03). In the bioassay, TSH/M22 induced luciferase-light was lower in CM from differentiated GO-OF than non-GO, suggesting that secreted TSHRv had neutralised their effects. α2 transcripts were present but reduced during adipogenesis (p < 0.005) with no difference observed between non-GO and GO. β5 transcripts were at the limit of detection. Our work demonstrated that TSHRv transcripts are expressed as protein, are more abundant in GO than non-GO OF and have the capacity to regulate signalling via the TSHR.
... Other studies have also shown an important role of PPARγ in the HA regulation of adipogenesis, albeit using cell lines [15,16]. HA may interact with its receptor CD44 to trigger IGF1 signalling [24], which is essential in adipogenesis through PPARγ regulation [2,25]. Furthermore, the chemical inhibition of HA production using 4-MU significantly enhanced adipogenesis, which is accompanied by increased PPARγ expression, ...
... Other studies have also shown an important role of PPARγ in the HA regulation of adipogenesis, albeit using cell lines [15,16]. HA may interact with its receptor CD44 to trigger IGF1 signalling [24], which is essential in adipogenesis through PPARγ regulation [2,25]. Furthermore, the chemical inhibition of HA production using 4-MU significantly enhanced adipogenesis, which is accompanied by increased PPARγ expression, particularly the adipose-specific isoform PPARγ2 [22]. ...
... The supernatant produced by centrifugation at 12,000× g provided the cytosolic fraction, and high salt extraction of the pellet to provide nuclear fraction. Nuclear protein samples were analysed by Western blotting using anti-PPARγ (from Cell Signalling Technology, Danvers, MA, USA) and anti-actin (from Santa Cruz Biotechnology, Dallas, Texas, USA) antibodies as previously described [25]. ...
Article
Full-text available
Hyaluronan (HA), an extra-cellular matrix glycosaminoglycan, may play a role in mesenchymal stem cell differentiation to fat but results using murine models and cell lines are conflicting. Our previous data, illustrating decreased HA production during human adipogenesis, suggested an inhibitory role. We have investigated the role of HA in adipogenesis and fat accumulation using human primary subcutaneous preadipocyte/fibroblasts (PFs, n = 12) and subjects of varying body mass index (BMI). The impact of HA on peroxisome proliferator-activated receptor gamma (PPARγ) expression was analysed following siRNA knockdown or HA synthase (HAS)1 and HAS2 overexpression. PFs were cultured in complete or adipogenic medium (ADM) with/without 4-methylumbelliferone (4-MU = HA synthesis inhibitor). Adipogenesis was evaluated using oil red O (ORO), counting adipogenic foci, and measurement of a terminal differentiation marker. Modulating HA production by HAS2 knockdown or overexpression increased (16%, p < 0.04) or decreased (30%, p = 0.01) PPARγ transcripts respectively. The inhibition of HA by 4-MU significantly enhanced ADM-induced adipogenesis with 1.52 ± 0.18- (ORO), 4.09 ± 0.63- (foci) and 2.6 ± 0.21-(marker)-fold increases compared with the controls, also increased PPARγ protein expression (40%, (p < 0.04)). In human subjects, circulating HA correlated negatively with BMI and triglycerides (r = −0.396 (p = 0.002), r = −0.269 (p = 0.038), respectively), confirming an inhibitory role of HA in human adipogenesis. Thus, enhancing HA action may provide a therapeutic target in obesity.
... Insulin, in common with insulin-like growth factor-1 (IGF-1) activates PI3 kinase (18) and MAP (mitogenactivated protein kinase) (19) pathways. Phosphorylation of protein kinase B (PKB/Akt) in turn phosphorylates forkhead box protein O1 (FOXO1) causing it to exit from the nucleus leading to increased transcription of adipogenic genes (20). Steroids induce the expression of the early adipogenic gene, CCAAT enhancer binding protein delta (C/EBP-d). ...
... Several studies, including from our group, have shown that activation of the TSHR in OF leads to an increase in hyaluronan production and adipogenesis (20,28). TSHR expression has been shown to increase during adipogenesis (5). ...
Article
Full-text available
Graves’ orbitopathy (GO) is a complex and poorly understood disease in which extensive remodeling of orbital tissue is dominated by adipogenesis and hyaluronan production. The resulting proptosis is disfiguring and underpins the majority of GO signs and symptoms. While there is strong evidence for the thyrotropin receptor (TSHR) being a thyroid/orbit shared autoantigen, the insulin-like growth factor 1 receptor (IGF1R) is also likely to play a key role in the disease. The pathogenesis of GO has been investigated extensively in the last decade with further understanding of some aspects of the disease. This is mainly derived by using in vitro and ex vivo analysis of the orbital tissues. Here, we have summarized the features of GO pathogenesis involving target autoantigens and their signaling pathways.
... 42 Therefore, the components of the GALGARO ET AL. Cell proliferation Bone marrow [11,18,42] Adipose tissue [43] Umbilical cord/blood [41,44] Amniotic fluid [45] Migration/Homing Bone marrow [46,47] Mesodermal differentiation potential Bone marrow [11,16,48] Amniotic fluid [45] Umbilical cord/blood [49] Other [50,51] Immunomodulation Bone marrow [19,20,47,[52][53][54][55][56][57][58][59][60][61][62] Adipose tissue [39,52,56,[63][64][65] Umbilical cord/blood [39,52,56,[65][66][67][68][69] Dental pulp/gingival [70][71][72][73][74][75][76][77][78][79] Other [39,[80][81][82][83][84] ADA Proliferation Bone marrow [85] Mesodermal differentiation potential Bone marrow [85] Others [51,[86][87][88][89] Immunomodulation Bone marrow [20,53] TNAP Mesodermal differentiation potential Bone marrow [90][91][92][93] Other [94][95][96] NPP1 Mesodermal differentiation potential Bone marrow [97] CD38 Proliferation Bone marrow [98] Migration/homing Bone marrow [98] Mesodermal differentiation potential Other [99,100] Immunomodulation Bone marrow [101] Amniotic membrane [102] Receptors A1R Proliferation Bone marrow [85] Adipose tissue [103] Dental pulp [103,104] Mesodermal differentiation potential Bone marrow [85,91,105] Adipose tissue [103] Dental pulp [103,104] Other [51,[106][107][108][109][110][111][112] adenosinergic pathway present in MSCs can give identity to different cell subpopulations and affect the physiological and pathological functions of these cells, as will be discussed in this review. ...
... 107 The activation of Gαs receptors inhibits the PPARγ transcription in a preadipocyte cell line (3T3-L1) and consequently the beginning of adipogenesis. 109 The pretreatment of preadipocytes or human BM-MSCs with the cAMP analog (8-Br-cAMP), or a cAMP agonist (forskolin), also reduced adipocytes differentiation. 110,133 Conversely, when 3T3-L1 cells were treated with specific adenylate cyclase inhibitor (SQ22536), an accelerated adipogenesis process occurred, with an increase of C/EBPα and PPARγ and enhanced Oil Red O staining. ...
Article
Full-text available
Mesenchymal stem cells (MSCs) play an important role in tissue homeostasis and damage repair through their ability to differentiate into cells of different tissues, trophic support, and immunomodulation. These properties made them attractive for clinical applications in regenerative medicine, immune disorders, and cell transplantation. However, despite multiple preclinical and clinical studies demonstrating beneficial effects of MSCs, their native identity and mechanisms of action remain inconclusive. Since its discovery, the CD73/ecto‐5′‐nucleotidase is known as a classic marker for MSCs, but its role goes far beyond a phenotypic characterization antigen. CD73 contributes to adenosine production, therefore, is an essential component of purinergic signaling, a pathway composed of different nucleotides and nucleosides, which concentrations are finely regulated by the ectoenzymes and receptors. Thus, purinergic signaling controls pathophysiological functions such as proliferation, migration, cell fate, and immune responses. Despite the remarkable progress already achieved in considering adenosinergic pathway as a therapeutic target in different pathologies, its role is not fully explored in the context of the therapeutic functions of MSCs. Therefore, in this review, we provide an overview of the role of CD73 and adenosine‐mediated signaling in the functions ascribed to MSCs, such as homing and proliferation, cell differentiation, and immunomodulation. Additionally, we will discuss the pathophysiological role of MSCs, via CD73 and adenosine, in different diseases, as well as in tumor development and progression. A better understanding of the adenosinergic pathway in the regulation of MSCs functions will help to provide improved therapeutic strategies applicable in regenerative medicine.
... In addition, it was postulated in the study that liver fibrosis-related proteins, namely alpha 2 macroglobulin and haptoglobin, were involved in the fibrosis of orbital tissues, hence progression of TED. Signalling cascades including PI3K-Akt and PPAR pathway, which were both previously documented as bearing substantial role in progression of TED [87,88], were shown to be upregulated in this study. ...
Article
Full-text available
Thyroid Eye Disease (TED) is an autoimmune disease that affects the extraocular muscles and periorbital fat. It most commonly occurs with Graves' Disease (GD) as an extrathyroidal manifestation, hence, it is also sometimes used interchangeably with Graves' Ophthalmopathy (GO). Well-known autoimmune markers for GD include thyroid stimulating hormone (TSH) receptor antibodies (TSH-R-Ab) which contribute to hyperthyroidism and ocular signs. Currently, apart from radiological investigations, detection of TED is based on clinical signs and symptoms which is largely subjective, with no established biomarkers which could differentiate TED from merely GD. We evaluated a total of 28 studies on potential biomarkers for diagnosis of TED. Articles included were published in English, which investigated clinical markers in tear fluid, orbital adipose-connective tissues, orbital fibroblasts and extraocular muscles, serum, thyroid tissue, as well as imaging biomark-ers. Results demonstrated that biomarkers with reported diagnostic power have high sensitivity and specificity for TED, including those using a combination of biomarkers to differentiate between TED and GD, as well as the use of magnetic resonance imaging (MRI). Other biomarkers which were upregulated include cytokines, proinflammatory markers, and acute phase reactants in subjects with TED, which are however, deemed less specific to TED. Further clinical investigations for these biomarkers, scrutinising their specificity and sensitivity on a larger sample of patients, may point towards selection of suitable biomarkers for aiding detection and prognosis of TED in the future.
... Confirming to our date Rocha et al. (2014) documented an age-related increase of PTEN in healthy rats. In addition, PI3K/AKT inhibition blocks adipogenic differentiation of brown preadipocytes (Fasshauer et al., 2000) and 3T3L1 preadipocytes (Zhang et al., 2009). Thus, the elevated PTEN protein we observed maybe one possible reason for the decrease in adipogenic differentiation in old VASCs. ...
Article
Full-text available
Stromal mesenchymal stem cells (MSCs) have the capability to self-renew and can differentiate into multiple cell types of the mesoderm germ layer, but their properties are affected by molecular aging mechanisms. MSCs can be obtained from adipose tissue termed as adipose-derived stem/stromal cells (ASCs) representing a promising tool for studying age-related diseases in detail. ASCs from young (16 weeks) and old (>108 weeks) rabbits were successfully isolated and propagated. ASCs showed the typical morphology and stained positive for CD105, Vimentin, Collagenase 1A, and negative for CD14, CD90, and CD73, demonstrating their mesenchymal origin. ASCs expressed MSC markers, including MYC, KLF4, CHD1, REST, and KAT6A, whereas pluripotency-related genes, such as NANOG, OCT4, and SOX2, were not expressed. Aged ASCs showed altered protein and mRNA levels of APOE, ATG7, FGF2, PTEN, and SIRT1. Adipogenic differentiation of old visceral ASCs was significantly decreased compared with young visceral ASCs. We successfully established rabbit ASC cultures representing an in vitro model for the analysis of stem cell aging mechanisms. ASCs, obtained from old female rabbits, showed age- and source-specific alteration due to aging of the donor. Stem cell plasticity was altered with age as shown by reduced adipogenic differentiation capacity.
... The PI3K-Akt pathway is activated by TSH-TSHR binding and its role in promoting orbital adipogenesis in GO patients has been described 59 . The peroxisome proliferator-activated receptor (PPAR) signaling pathway was associated with thyroid-eye disease progression 60 . Several pathways associated with hypertrophic cardiomyopathy were also identified, which may be associated with the additional strain imposed on the cardio-circulatory system by hyperthyroidism. ...
Article
Full-text available
Graves' Disease (GD) is an autoimmune condition in which thyroid-stimulating antibodies (TRAB) mimic thyroid-stimulating hormone function causing hyperthyroidism. 5% of GD patients develop inflammatory Graves' orbitopathy (GO) characterized by proptosis and attendant sight problems. A major challenge is to identify which GD patients are most likely to develop GO and has relied on TRAB measurement. We screened sera/plasma from 14 GD, 19 GO and 13 healthy controls using high-throughput proteomics and miRNA sequencing (Illumina's HiSeq2000 and Agilent-6550 Funnel quadrupole-time-of-flight mass spectrometry) to identify potential biomarkers for diagnosis or prognosis evaluation. Euclidean distances and differential expression (DE) based on miRNA and protein quantification were analysed by multidimensional scaling (MDS) and multinomial regression respectively. We detected 3025 miRNAs and 1886 proteins and MDS revealed good separation of the 3 groups. Biomarkers were identified by combined DE and Lasso-penalized predictive models; accuracy of predictions was 0.86 (±0:18), and 5 miRNA and 20 proteins were found including Zonulin, Alpha-2 macroglobulin, Beta-2 glycoprotein 1 and Fibronectin. Functional analysis identified relevant metabolic pathways, including hippo signaling, bacterial invasion of epithelial cells and mRNA surveillance. Proteomic and miRNA analyses, combined with robust bioinformatics, identified circulating biomarkers applicable to diagnose GD, predict GO disease status and optimize patient management.
... Target gene mRNA expression was normalized to acidic ribosomal protein (Rplp0) as a reference gene. MCAM and Perilipin 1 (Plin1) primers were designed using Primer3 v0.4.0 software [20], Rplp0, Peroxisome proliferator-activated receptor gamma isoform 2 (PPARγ2), and Adiponectin (Adipoq) primer sequences were obtained from published reports [21][22][23]. The sequences of primers used are shown in Table 1. ...
Article
Full-text available
We investigated for the first time the expression of melanoma cell adhesion molecule (MCAM) and its involvement in the differentiation of 3T3-L1 fibroblasts to adipocytes. We found that MCAM mRNA increased subsequent to the activation of the master regulator of adipogenesis, PPARγ, and this increase was maintained in the mature adipocytes. On the other hand, MCAM knockdown impaired differentiation and induction of PPARγ as well as expression of genes activated by PPARγ. However, events that precede and are necessary for early PPARγ activation, such as C/EBPβ induction, β-catenin downregulation, and ERK activation, were not affected in the MCAM knockdown cells. In keeping with this, the increase in PPARγ mRNA that precedes MCAM induction was not altered in the knockdown cells. In conclusion, our findings suggest that MCAM is a gene upregulated and involved in maintaining PPARγ induction in the late but not in the early stages of 3T3-L1 fibroblasts adipogenesis.
Article
Full-text available
Purpose: Graves' orbitopathy (GO) is an orbital manifestation of autoimmune Graves' disease, and orbital fibroblast is considered a target cell, producing pro-inflammatory cytokines and/or differentiating into adipocytes. Adipose tissue has been focused on as an endocrine and inflammatory organ secreting adipokines. We investigated the pathogenic role of a specific adipokine, adipsin, known as complement factor D in Graves' orbital fibroblasts. Methods: The messenger RNA (mRNA) expression of multiple adipokines was investigated in adipose tissues harvested from GO and healthy subjects. Adipsin protein production was analyzed in primary cultured orbital fibroblasts under insulin growth factor (IGF)-1, CD40 ligand (CD40L) stimulation, and adipogenesis. The effect of blocking adipsin with small interfering RNA (siRNA) on pro-inflammatory cytokine production and adipogenesis was evaluated using quantitative real-time PCR, Western blot, and ELISA. Adipogenic differentiation was identified using Oil Red O staining. Results: Adipsin gene expression was significantly elevated in GO tissue and increased after the stimulation of IGF-1 and CD40L, as well as adipocyte differentiation in GO cells. Silencing of adipsin suppressed IGF-1-induced IL-6, IL-8, COX2, ICAM-1, CCL2 gene expression, and IL-6 protein secretion. Adipsin suppression also attenuated adipocyte differentiation. Exogenous treatment of recombinant adipsin resulted in the activation of the Akt, ERK, p-38, and JNK signaling pathways. Conclusions: Adipsin, secreted by orbital fibroblasts, may play a distinct role in the pathogenesis of GO. Inhibition of adipsin ameliorated the production of pro-inflammatory cytokines and adipogenesis in orbital fibroblasts. Our study provides an in vitro basis suggesting adipsin as a potential therapeutic target for GO treatment.
Article
Full-text available
Introduction Graves’ disease (GD) is an autoimmune disorder caused by autoantibodies against the thyroid stimulating hormone receptor (TSHR) leading to overstimulation of the thyroid gland. Thyroid eye disease (TED) is the most common extra thyroidal manifestation of GD. Therapeutic options to treat TED are very limited and novel treatments need to be developed. In the present study we investigated the effect of linsitinib, a dual small-molecule kinase inhibitor of the insulin-like growth factor 1 receptor (IGF-1R) and the Insulin receptor (IR) on the disease outcome of GD and TED. Methods Linsitinib was administered orally for four weeks with therapy initiating in either the early (“active”) or the late (“chronic”) phases of the disease. In the thyroid and the orbit, autoimmune hyperthyroidism and orbitopathy were analyzed serologically (total anti-TSHR binding antibodies, stimulating anti TSHR antibodies, total T4 levels), immunohistochemically (H&E-, CD3-, TNFa- and Sirius red staining) and with immunofluorescence (F4/80 staining). An MRI was performed to quantify in vivo tissue remodeling inside the orbit. Results Linsitinib prevented autoimmune hyperthyroidism in the early state of the disease, by reducing morphological changes indicative for hyperthyroidism and blocking T-cell infiltration, visualized by CD3 staining. In the late state of the disease linsitinib had its main effect in the orbit. Linsitinib reduced immune infiltration of T-cells (CD3 staining) and macrophages (F4/80 and TNFa staining) in the orbita in experimental GD suggesting an additional, direct effect of linsitinib on the autoimmune response. In addition, treatment with linsitinib normalized the amount of brown adipose tissue in both the early and late group. An in vivo MRI of the late group was performed and revealed a marked decrease of inflammation, visualized by ¹⁹ F MR imaging, significant reduction of existing muscle edema and formation of brown adipose tissue. Conclusion Here, we demonstrate that linsitinib effectively prevents development and progression of thyroid eye disease in an experimental murine model for Graves’ disease. Linsitinib improved the total disease outcome, indicating the clinical significance of the findings and providing a path to therapeutic intervention of Graves’ Disease. Our data support the use of linsitinib as a novel treatment for thyroid eye disease.
Article
Full-text available
Adipocyte differentiation is driven by the expression and activation of three transcription factor families: the differentially expressed CAAT/enhancer binding proteins (C/EBPs) alpha, beta, and delta; the helix-loop-helix adipocyte differentiation and determination factor-1; and peroxisome proliferator activated receptor gamma (PPAR gamma), expressed as two isoforms, PPAR gamma 1 and the adipocyte-specific PPAR gamma 2, Overexpression of PPAR gamma can induce adipocyte differentiation; therefore, we analyzed the expression of the two PPAR gamma isoforms during early stages of differentiation to determine whether one was preferentially induced as an early determining event, Surprisingly, in the first 24 h, a 3-6-fold increase of PPAR gamma 2 mRNA was observed, whereas PPAR gamma 1 mRNA remained unchanged. PPAR gamma 1 was induced 1 day later. Overexpression of C/EBP beta has also been shown to induce adipocyte differentiation. A C/EBP site was identified only in the human PPAR gamma 2 promoter. Its deletion blunted the response of PPAR gamma 2 promoter to cotransfected C/EBP beta or methylisobutylxanthine treatment. We hypothesize that PPAR gamma 2 initiates adipocyte differentiation.
Article
Full-text available
Gsα regulates the differentiation of 3T3-L1 mouse embryonic fibroblasts to adipocytes, a process termed adipogenesis. Inducers of adipogenesis lead to a loss of Gsα and derepress differentiation to adipocytes. The broad spectrum tyrosine kinase inhibitor genistein is shown to block induction of adipogenesis, suggesting an early role of tyrosine phosphorylation in adipogenesis. Staining of phosphotyrosine identified prominent staining of a ∼70-kDa protein, hypothesized to be the tyrosine kinase Syk. Reverse transcription and polymerase chain reaction amplification established the expression of Syk mRNA in these embryonic fibroblasts. Immunoprecipitations with Syk-specific antibodies demonstrated the presence of Syk in fibroblasts and a rapid increase in the amount of phospho-Syk, peaking at 24 h post induction. Clones constitutively expressing Gsα, which can no longer be induced to differentiate, no longer display increased phospho-Syk levels in response to inducers. The linkage between Gsα and Syk was probed by immunoprecipitations revealing association of Syk with Gsα in the absence of induction. Upon induction of adipogenesis, Gsα levels decline and phospho-Syk levels as well as Syk kinase activity increase. Expression of wild-type Syk both potentiates the ability of inducers to act as well as induces adipogenesis itself. Expression of the kinase-deficient Syk had no such effects on adipogenesis. These data provide a new insight into the control of adipogenesis, suggesting that Gsα represses adipogenesis via Syk. Treatment with the inducers promotes a decline in Gsα, increases in levels of phospho-Syk, and adipogenesis.
Article
Full-text available
Both the α and βγ subunits of heterotrimeric guanine nucleotide–binding proteins (G proteins) communicate signals from receptors to effectors. Gβγ subunits can regulate a diverse array of effectors, including ion channels and enzymes. Gα subunits bound to guanine diphosphate (Gα-GDP) inhibit signal transduction through Gβγ subunits, suggesting a common interface on Gβγ subunits for Gα binding and effector interaction. The molecular basis for interaction of Gβγ with effectors was characterized by mutational analysis of Gβ residues that make contact with Gα-GDP. Analysis of the ability of these mutants to regulate the activity of calcium and potassium channels, adenylyl cyclase 2, phospholipase C-β2, and β-adrenergic receptor kinase revealed the Gβ residues required for activation of each effector and provides evidence for partially overlapping domains on Gβ for regulation of these effectors. This organization of interaction regions on Gβ for different effectors and Gα explains why subunit dissociation is crucial for signal transmission through Gβγ subunits.
Article
Full-text available
Adipocyte differentiation is an important component of obesity and other metabolic diseases. This process is strongly inhibited by many mitogens and oncogenes. Several growth factors that inhibit fat cell differentiation caused mitogen-activated protein (MAP) kinase-mediated phosphorylation of the dominant adipogenic transcription factor peroxisome proliferator-activated receptor γ (PPARγ) and reduction of its transcriptional activity. Expression of PPARγ with a nonphosphorylatable mutation at this site (serine-112) yielded cells with increased sensitivity to ligand-induced adipogenesis and resistance to inhibition of differentiation by mitogens. These results indicate that covalent modification of PPARγ by serum and growth factors is a major regulator of the balance between cell growth and differentiation in the adipose cell lineage.
Article
Full-text available
Phosphoenolpyruvate carboxykinase (PEPCK) is expressed at high levels in liver, kidney, and adipose tissue. This enzyme catalyzes the rate-limiting step in hepatic and renal gluconeogenesis and adipose glyceroneogenesis. The regulatory factors important for adipose expression of the PEPCK gene are not well defined. Previous studies with transgenic mice established that the region between bp -2086 and -888 is required for expression in adipose tissue but not for expression in liver or kidney tissue. We show here that a DNA fragment containing this region can function as an enhancer and direct differentiation-dependent expression of a chloramphenicol acetyltransferase gene from a heterologous promoter in cultured 3T3-F442A preadipocytes and adipocytes. We further demonstrate that the adipocyte-specific transcription factor PPAR gamma 2, previously identified as a regulator of the adipocyte P2 enhancer, binds in a heterodimeric complex with RXR alpha to the PEPCK 5'-flanking region at two sites, termed PCK1 (bp -451 to -439) and PCK2 (bp -999 to -987). Forced expression of PPAR gamma 2 and RXR alpha activates the PEPCK enhancer in non-adipose cells. This activation is potentiated by peroxisome proliferators and fatty acids but not by 9-cis retinoic acid. Mutation of the PPAR gamma 2 binding site (PCK2) abolishes both the activity of the enhancer in adipocytes and its ability to be activated by PPAR gamma 2 and RXR alpha. These results establish a role for PPAR gamma 2 in the adipose expression of the PEPCK gene and suggest that this factor functions as a coordinate regulator of multiple adipocyte-specific genes.
Article
Fully-differentiated mouse 3T3-L1 fibroblasts accumulate large amounts of lipid at 7-10 days after induction by insulin or by dexamethasone and a methyl xanthine. G proteins mediate transmembrane signalling from a diverse group of cell-surface receptors to effector units that include phospholipase C, adenylyl cyclase and ion channels. They are also targets of regulation themselves. 3T3-L1 fibroblasts display marked changes in levels of G protein when induced to differentiate to adipocytes. Here we show that cholera toxin, which ADP-ribosylates and activates the G protein subunit Gs alpha, blocks the induction of differentiation, whereas increasing intracellular cyclic AMP directly with the dibutyryl analogue or indirectly with pertussis toxin or forskolin does not affect differentiation. Oligodeoxynucleotides antisense to the sequence encoding Gs alpha accelerate differentiation markedly. The time course of adipogenesis declined from 7-10 days in controls to roughly 3 days in cultures treated with antisense-Gs alpha oligodeoxynucleotides, whereas oligodeoxynucleotides, antisense to Gi alpha 1, Gi alpha 3, and sense and missense to Gs alpha, had no such effect. Antisense-Gs alpha alone induced differentiation by day 7, indicating that Gs alpha activity modulates differentiation in 3T3-L1 cells, acting in a new role which is independent of increased intracellular cAMP.
Article
I. Introduction IN classical physiology, the main regulation of the thyroid gland involves a positive control by pituitary TSH. The pituitary thyrotrophs that synthesize and secrete TSH are submitted to tonic stimulation by the hypothalamic TRH and to negative feedback by T3, which is generated in these cells from plasma T4 (1, 2). The predominant role of TSH in thyroid regulation is exemplified by physiological experiments and pathological situations. Hypophysectomy, or treatment with thyroid hormones that reduces plasma TSH levels without interfering with other pituitary functions, almost abolishes thyroid function and greatly reduces the weight and cell mass of the gland. Conversely, chronic inhibition of the synthesis of thyroid hormones, and thus of their secretion, which greatly enhances TSH plasma levels, enhances thyroid growth and iodine metabolism upstream from the inhibited step. In pathology, hyperthyrotropinemia caused by pituitary adenomas induces hyperthyroidism and goiter. Hypopituitarism ...
Article
The original Whickham Survey documented the prevalence of thyroid disorders in a randomly selected sample of 2779 adults which matched the population of Great Britain in age, sex and social class. The aim of the twenty-year follow-up survey was to determine the incidence and natural history of thyroid disease in this cohort. Subjects were traced at follow-up via the Electoral Register, General Practice registers, Gateshead Family Health Services Authority register and Office of Population Censuses and Surveys. Eight hundred and twenty-five subjects (30% of the sample) had died and, in addition to death certificates, two-thirds had information from either hospital/General Practitioner notes or post-mortem reports to document morbidity prior to death. Of the 1877 known survivors, 96% participated in the follow-up study and 91% were tested for clinical, biochemical and immunological evidence of thyroid dysfunction. Outcomes in terms of morbidity and mortality were determined for over 97% of the original sample. The mean incidence (with 95% confidence intervals) of spontaneous hypothyroidism in women was 3.5/1000 survivors/year (2.8-4.5) rising to 4.1/1000 survivors/year (3.3-5.0) for all causes of hypothyroidism and in men was 0.6/1000 survivors/year (0.3-1.2). The mean incidence of hyperthyroidism in women was 0.8/1000 survivors/year (0.5-1.4) and was negligible in men. Similar incidence rates were calculated for the deceased subjects. An estimate of the probability of the development of hypothyroidism and hyperthyroidism at a particular time, i.e. the hazard rate, showed an increase with age in hypothyroidism but no age relation in hyperthyroidism. The frequency of goitre decreased with age with 10% of women and 2% of men having a goitre at follow-up, as compared to 23% and 5% in the same subjects respectively at the first survey. The presence of a goitre at either survey was not associated with any clinical or biochemical evidence of thyroid dysfunction. In women, an association was found between the development of a goitre and thyroid-antibody status at follow-up, but not initially. The risk of having developed hypothyroidism at follow-up was examined with respect to risk factors identified at first survey. The odds ratios (with 95% confidence intervals) of developing hypothyroidism with (a) raised serum TSH alone were 8 (3-20) for women and 44 (19-104) for men; (b) positive anti-thyroid antibodies alone were 8 (5-15) for women and 25 (10-63) for men; (c) both raised serum TSH and positive anti-thyroid antibodies were 38 (22-65) for women and 173 (81-370) for men. A logit model indicated that increasing values of serum TSH above 2mU/l at first survey increased the probability of developing hypothyroidism which was further increased in the presence of anti-thyroid antibodies. Neither a positive family history of any form of thyroid disease nor parity of women at first survey was associated with increased risk of developing hypothyroidism. Fasting cholesterol and triglyceride levels at first survey when corrected for age showed no association with the development of hypothyroidism in women. This historical cohort study has provided incidence data for thyroid disease over a twenty-year period for a representative cross-sectional sample of the population, and has allowed the determination of the importance of prognostic risk factors for thyroid disease identified twenty years earlier.