ArticlePDF Available

Inflammation and insulin resistance exert dual effects on adipose tissue tumor protein 53 expression

Authors:

Abstract and Figures

Objective: The purpose of this study was to investigate the expression of human adipose tissue protein 53 (p53) in subjects who varied widely in terms of obesity and insulin resistance. We also analyzed different in vivo and in vitro models to try to comprehend the associations found in humans. Methods: p53 was analyzed in human adipose and isolated adipocytes, in high fat-fed and GLP-1R KO mice, during in vitro adipogenesis, and in adipocytes after high glucose, rosiglitazone and inflammatory conditions. The effects of surgery-induced weight loss and ex vivo metformin were also evaluated. Results: Omental (OM) p53 gene expression (+27%, P=0.001) and protein (+11%, P=0.04) were increased in obese subjects and high fat diet-induced obese mice (+86%, P=0.018). Although the obesity-associated inflammatory milieu was associated with increased OM p53, this was negatively related to insulin resistance and glycated hemoglobin, and positively with biomarkers for insulin sensitivity. Multiple linear regression analyses revealed that glycated hemoglobin (P<0.0001) and body mass index (P=0.048) contributed independently to explain 13.7% (P<0.0001) of the OM p53 variance. Accordingly, the improvement of insulin sensitivity with surgery-induced weight loss (+51%, P=0.01) and metformin (+42%, P=0.02) led to increased adipose p53. While the glucose-intolerant GLP-1R KO mice showed decreased mesenteric p53 (-45.4%, P=0.017), high glucose led to decreased p53 in pre-adipocytes (-27%, P<0.0001). Inflammatory treatments led to increased p53 (+35%, P<0.0001), while Rs downregulated this expression (-40%, P=0.005) in mature adipocytes. Conclusion: Inflammation and insulin resistance exert dual effects on adipose p53, which seems to be the final result of these opposing forces.
Content may be subject to copyright.
ORIGINAL ARTICLE
Inflammation and insulin resistance exert dual effects on adipose
tissue tumor protein 53 expression
FJ Ortega
1,2
, JM Moreno-Navarrete
1,2
, D Mayas
1,3
, M Serino
4
, JI Rodriguez-Hermosa
5
, W Ricart
1,2
, E Luche
4
, R Burcelin
4
, FJ Tinahones
1,3
,
GFru¨ hbeck
1,6
, G Mingrone
7
and JM Ferna´ndez-Real
1,2
OBJECTIVE: The purpose of this study was to investigate the expression of human adipose tissue protein 53 (p53) in subjects who
varied widely in terms of obesity and insulin resistance. We also analyzed different in vivo and in vitro models to try to comprehend
the associations found in humans.
METHODS: p53 was analyzed in human adipose and isolated adipocytes, in high fat-fed and GLP-1R KO mice, during in vitro
adipogenesis, and in adipocytes after high glucose, rosiglitazone and inflammatory conditions. The effects of surgery-induced
weight loss and ex vivo metformin were also evaluated.
RESULTS: Omental (OM) p53 gene expression ( þ 27%, P ¼ 0.001) and protein ( þ 11%, P ¼ 0.04) were increased in obese subjects
and high fat diet-induced obese mice ( þ 86%, P ¼ 0.018). Although the obesity-associated inflammatory milieu was associated with
increased OM p53, this was negatively related to insulin resistance and glycated hemoglobin, and positively with biomarkers for
insulin sensitivity. Multiple linear regression analyses revealed that glycated hemoglobin (Po0.0001) and body mass index
(P ¼ 0.048) contributed independently to explain 13.7% (Po0.0001) of the OM p53 variance. Accordingly, the improvement of
insulin sensitivity with surgery-induced weight loss ( þ 51%, P ¼ 0.01) and metformin ( þ 42%, P ¼ 0.02) led to increased adipose
p53. While the glucose-intolerant GLP-1R KO mice showed decreased mesenteric p53 ( 45.4%, P ¼ 0.017), high glucose led to
decreased p53 in pre-adipocytes ( 27%, Po0.0001). Inflammatory treatments led to increased p53 ( þ 35%, Po0.0001), while
Rs downregulated this expression ( 40%, P ¼ 0.005) in mature adipocytes.
CONCLUSION: Inflammation and insulin resistance exert dual effects on adipose p53, which seems to be the final result of these
opposing forces.
International Journal of Obesity advance online publication, 1 October 2013; doi:10.1038/ijo.2013.163
Keywords: tumor protein 53; adipocytes; adipose tissue; inflammation; insulin resistance; type 2 diabetes
INTRODUCTION
The tumor suppressor activity of the protein 53 (p53) has been
clearly established in the last three decades (see Levine and Oren
1
for a review). Less attention has been given to its possible
involvement in other biological processes such as cell metabolism
and development. Notably, inflammatory stress triggers the p53
response,
2
and p53 itself is also emerging as an important
regulator of metabolic homeostasis.
3
Both adipose tissue and systemic inflammation are up-
regulated in obese subjects, having a crucial role in the
progression of metabolic disturbances.
4,5
Omental (OM) adipose
tissue dysfunction is known to act as a source of inflammatory-
related factors in individuals with central obesity.
6
On the other
hand, adipose tissue senescence has been recently recognized to
be linked to inflammation in obese subjects with shortened
telomeres.
5,7
Minamino et al.
8
recently reported that p53 in mice adipose
tissue was closely associated with cellular aging, a new target for
the treatment of diabetes. These authors used Ay mice, known to
develop obesity and diabetes in the context of increased oxidative
stress and pro-inflammatory cytokines.
8
They demonstrated the
induction of senescence-like changes (for example, increased p53
and shortened telomeres) in adipose tissue from these mice.
8
Results from this study suggested that p53 in adipose tissue may
lead to increased oxidative stress-induced nuclear transcription
factor (NF)-kB activation and the induction of inflammatory
cytokines, promoting at the end the development of insulin
resistance in mice.
Recent findings also demonstrate that p53 modulates the
coordinated interplay between proliferation and differentiation in
many cell lineages,
9–11
including adipocytes,
12
having unsuspected
effects on cellular metabolism.
13,14
Molchadsky et al.
15
demonstrated
that p53 has opposing roles in a cell fate-dependent manner,
inhibiting adipogenesis but promoting muscle differentiation. In vitro
and in vivo studies have also provided interesting insights into the
association of p53 with adipocytes’ fate decisions (see Bazuine et al.
12
for a review), its involvement in adipose tissue dysfunction
8
and in
metabolic homeostasis.
1
CIBER de la Fisiopatologı
´
a de la Obesidad y Nutricio
´
n (CIBERobn, CB06/03/0010), and Instituto de Salud Carlos III (ISCIII), Santiago de Compostela, Spain;
2
Service of Diabetes,
Endocrinology and Nutrition (UDEN), Institut d’Investigacio
´
Biome
`
dica de Girona (IdIBGi), Hospital ‘‘Dr Josep Trueta’’ of Girona, Girona, Spain;
3
Service of Endocrinology and
Nutrition, Hospital Clı
´
nico Universitario Virgen de Victoria de Ma
´
laga, Ma
´
laga, Spain;
4
INSERM Unite
´
858, Institut de Me
´
decine Mole
´
culaire de Rangueil, Universite
´
Paul Sabatier,
IFR31, Toulouse, France;
5
Department of Surgery, Institut d’Investigacio
´
Biome
`
dica de Girona (IdIBGi), Girona, Spain;
6
Department of Endocrinology & Nutrition, Clı
´
nica
Universidad de Navarra–CIBERobn CBO6/03/1014, Pamplona, Spain and
7
Institute of Internal Medicine, Catholic University of Rome, Rome, Italy. Correspondence:
Dr JM Ferna
´
ndez-Real or Dr FJ Ortega, Service of Diabetes, Endocrinology and Nutrition (UDEN), Institut d’Investigacio
´
Biome
`
dica de Girona (IdIBGi), Hospital ‘‘Dr Josep Trueta’’ of
Girona, Carretera de Franc¸a s/n, Girona 17007, Spain.
E-mail: jmfreal@idibgi.org or fortega@idibgi.org
Received 21 March 2013; revised 2 July 2013; accepted 5 August 2013; accepted article preview online 3 September 2013
International Journal of Obesity (2013), 19
&
2013 Macmillan Publishers Limited All rights reserved 0307-0565/13
www.nature.com/ijo
In fact, p53 is able to contribute to the regulation of many
cellular pathways, including glycolysis,
16
insulin sensitivity,
14
fatty
acid oxidation and mTOR signaling.
17
Overall, the ability of p53 to
respond to nutrient deficiencies in insulin-sensitive tissues is
consistent with its established function as a mediator of stress,
3
but p53 seems to have many pleiotropic effects,
16,18
which remain
to be elucidated in this context.
The challenge is to define the specific role of adipose tissue p53
in association with inflammation and the development of obesity
and obesity-related complications, such as insulin resistance and
type 2 diabetes. Very few data are available regarding the
expression of p53 and its modulating partner, murine double
minute 2 (MDM2, a proto-oncogene that is closely associated with
p53 gene products), in human adipose tissue. Thus, the purpose of
this study was to investigate the expression of p53 and MDM2 in
subcutaneous (SC) and OM adipose tissue from a large cohort of
subjects who varied widely in terms of obesity and insulin
resistance. We also analyzed different in vivo and in vitro models to
try to comprehend the associations found in human adipose
tissue.
MATERIALS AND METHODS
Human studies
Subjects and samples. A total of 193 OM and 113 SC adipose tissue
samples (76 paired fat samples) were obtained from human fat depots
during elective surgical procedures. These fat samples were provided from
a group of 230 subjects (76 men and 154 women) with a body mass index
between 18 and 70 kg m
2
who were invited to participate at the
Endocrinology Service of the Hospital Universitari de Girona Dr Josep
Trueta (Girona, Spain) and the Hospital Carlos Haya de Ma
´
laga (Ma
´
laga,
Spain). All subjects were of Caucasian origin and reported that their body
weight had been stable for at least 3 months before the study. They had
no systemic disease other than type 2 diabetes and/or obesity, and all were
free of any infections within the previous months before the study. All
subjects gave written informed consent after the purpose of the study was
explained to them. Approximately 5 g of SC and OM adipose tissue fresh
samples from 12 subjects (24 paired samples) was used to isolate stromal-
vascular cells and MAs, as previously described.
19
Anthropometric and analytical measurements. BMI was calculated as
weight (in kg) divided by height (in m) squared. Bioelectric impedance
and/or air-displacement plethysmoghraphy were used to estimate the
body fat composition in those subjects. According to these anthropometric
parameters, subjects were classified as non-obese (BMIo30.0 kg m
2
) and
obese (BMIX30.0 kg m
2
, and body fat % X25% for men and X35% for
women) subjects. Data from a standard 75-g oral glucose tolerance test,
performed after an overnight fast, and venous blood samples that were
drawn at time points zero, 30, 60, 90 and 120 min for the determination of
plasma glucose concentrations, were available for a random subpopulation
of 28 participants (8 men and 20 women). Area under the curve (AUC) of
glucose concentrations during the entire 120 min of the 75-g oral glucose
tolerance test was calculated according to the trapezoid method as:
0.5 (0.5 c0 þ c30 þ c60 þ c90 þ 0.5 c120), where c is the concentration.
Other measurements were performed using the usual techniques of the
clinical laboratory.
Study of the effects of weight loss induced by bariatric surgery
Seven Caucasian morbidly obese (BMI ¼ 50.4
±
9.0 kg m
2
, age ¼ 40
±
10
years (mean
±
s.d.)) and normotolerant women were recruited. An oral
glucose tolerance test, an intravenous glucose tolerance test and a
euglycemic hyperinsulinemic clamp were randomly performed within
1 month before surgery and 1 month after surgery. The main procedures
were as previously described.
20
All subjects were nonsmokers and were
not receiving statins or antidiabetic medication. Patients with signs of
infection were excluded. The study protocol was approved by the
institutional ethics committee of the Catholic University of Rome (Rome,
Italy). The nature and purpose of the study were carefully explained to all
the subjects before they provided their written consent to participate.
Ex vivo studies in human adipose tissue
Explants of SC adipose tissue were obtained from five obese female
participants undergoing elective open abdominal surgery (gastrointestinal
bypass) after an overnight fast, cultured and treated as previously
described.
21
The mean age was 46
±
6.4 years (range, 39–58 years) and
the BMI 44.9
±
12.4 kg m
2
. None of the subjects had a history of hepatic
or renal disorders. The study had the approval of the ethics committee of
the participant institutions and all patients gave informed written consent.
Mice model experiments
Experimental protocols were conducted in accordance with the French
government policies and were validated by the local animal ethics
committee of the Rangueil Hospital and the Services ve
´
terinaires de la
Sante
´
et de la production animale, Ministe
`
re franc¸ais de lervices ve
´
te.
C57bl6 adult male mice were housed for less than 12 h of light/12 h of dark
cycle, in a temperature-controlled environment. Body weight and food
intake were measured weekly. In all the experiments, mice were killed after
overnight fasting. White mesenteric fat, taken as a reflection of visceral
adiposity, and/or SC adipose tissue samples were then removed, weighed
and either fixed in buffered formalin or snap frozen in liquid nitrogen. All
samples were stored at 80 1C until use.
p53 expression in mesenteric fat from GLP-1R KO mice and high-fat diet-fed
mice. Four- to five-week-old C57BL/6J male mice (Iffa-Credo, L’Arbresle,
France) were fed with normal chow (energy content: 12% fat, 28% protein
and 60% carbohydrate) or a high-fat carbohydrate-free diet (energy
content: 72% fat (corn oil and lard), 28% protein and o1% carbohydrate
22
)
for 4 weeks (acute stimuli) or 9 months (diet-induced obese mice). Notably,
this diet has been defined to promote a heterogeneous obesity status.
Hence, it is suitable to study the natural and diverse adaptation of mice
to metabolic stress, and to provide a heterogeneous population that can
be secondarily classified in obese and non-obese mice, as previously
described.
23
In vitro studies
Culture and differentiation of human adipocytes. Isolated pre-adipocytes
(Zen-Bio Inc., Research Triangle Park, NC, USA) were plated and
differentiated to MAs as previously described.
24
Pre-adipocytes and MAs
were harvested and stored at 80 1C for RNA extraction to study gene
expression levels during adipogenesis. The experiment was performed in
triplicate for each sample.
Effects of MCM and Rs on p53 gene expression. Adipocytes were incubated
with fresh media (control), fresh media containing macrophage-condi-
tioned medium (MCM, 5%) from THP-1 cells previously treated with
10 ng ml
1
lipopolysaccharides (LPS) or fresh media containing Rs (2 mM).
After 48 h, the supernatants were centrifuged at 400 g for 5 min, the cells
harvested, and pellets and supernatants were stored at 80 1C for RNA
and protein analysis.
The embryonic fibroblast mouse cell line 3T3-L1 (Zen-Bio Inc.) was
maintained in Dulbecco’s modified Eagle’s medium containing 20 m
M of
glucose, 10% fetal bovine serum, 100 units ml
1
penicillin and
100 mgml
1
streptomycin. Two days after confluence, insulin (5 mgml
1
),
dexamethasone (0.5 m
M) and IBMX (0.5 mM) mixture was added and
maintained for 2 days, followed by 5 days with a medium containing
insulin (5 mgml
1
). Cells were then considered MAs, harvested and stored
at 80 1C. In parallel, non-differentiated 3T3-L1 were maintained for
7 days in Dulbecco’s modified Eagle’s medium containing 10 (low),
20 (normal) and 200 (high) m
M of glucose, 10% fetal bovine serum,
100 units ml
1
penicillin and 100 mgml
1
streptomycin to analyze the
specific effect of glucose on p53 gene expression.
Gene expression analyses
RNA was prepared from both fat biopsy samples and cellular debris using
the RNeasy Lipid Tissue Mini Kit (Qiagen, Gaithersburg, MD, USA). The
integrity of each RNA sample was checked with an Agilent Bioanalyzer
(Agilent Technologies, Palo Alto, CA, USA). Total RNA was quantified by
means of the GeneQuant spectrophotometer (GE Health Care, Piscataway,
NJ, USA) and 3 mg of RNA was then reverse transcribed to cDNA using the
High Capacity cDNA Archive Kit (Applied Biosystems, Darmstadt, Germany)
according to the manufacturer’s protocol.
Gene expression was assessed by real-time PCR using the LightCycler
480 Real-Time PCR System (Roche Diagnostics, Barcelona, Spain), using
p53 in human adipose tissue
FJ Ortega et al
2
International Journal of Obesity (2013) 1 9 & 2013 Macmillan Publishers Limited
TaqMan technology suitable for relative gene expression quantification.
The reaction was performed following the manufacturer’s protocol in a
final volume of 7 ml. The cycle program consisted of an initial denaturing of
10 min at 95 1C, then 45 cycles of a 15-s denaturizing phase at 92 1C and a
1-min annealing and extension phase at 60 1C. Replicates and positive and
negative controls were included in all the reactions. The commercially
available and pre-validated TaqMan primer/probe sets used were as
follows: Cyclophilin A (PPIA, 4333763) was used as endogenous control
for all target genes in each reaction, and tumor-related protein 53
(p53, Hs01034249_m1), transformed mouse 3T3 cell double minute 2
(MDM2, Hs00234753_m1), solute carrier family 2 (facilitated glucose
transporter) member 4 (GLUT4, Hs00168966_m1), insulin receptor
substrate 1 (IRS1, Hs00178563_m1), leptin (LEP, Hs00174877_m1), fatty
acid synthase (FASN, Hs00188012_m1), tumor necrosis factor-alpha
(TNF-a, Hs01113624_g1), interleukin-6 (IL-6, Hs00985639_m1), adiponectin
(ADIPOQ, Hs00605917_m1) and the cluster differentiation molecules 14
(CD14, Hs02621496_s1) and 68 (CD68, Hs02836816_g1) were the target
genes. P53 (Mm01731287_m1), calgranulin A (s100a8, Mm00496696_g1),
lipopolysaccharide-binding protein (lbp, Mm00493139_m1) and fatty acid
binding protein 4 (fabp4, Mm00445878_m1) gene expression in mice was
assessed and expressed as a ratio relative to eukaryotic 18S rRNA
(18S, Hs99999901_s1). The second derivative maximum method was used
for the determination of the crossing points (Cps). A Cp value was obtained
for each amplification curve, and the DCp value was first calculated by
subtracting the Cp value for human PPIA or 18S from the Cp value for each
sample and transcript. Fold changes compared with the endogenous
control were then determined by calculating 2
Df
, so gene expression
results are expressed in all cases as an expression ratio relative to the
endogenous control expression, according to the manufacturer’s
instructions.
p53 protein analyses by ELISA
Trying to evaluate protein levels and the activation degree of p53 in
human adipose tissue (n ¼ 28; OM and SC paired fat samples), we used
three solid-phase sandwich enzyme-linked immunosorbent assays
(ELISAs) that detect endogenous levels of total p53 and phospho-p53
(Ser15) protein, and acetylated lysines in p53, respectively (Cell
Signaling, Izasa S.A., Barcelona, Spain). The analyses were performed
following the manufacturer’s instructions. After incubation with cell
lysates, total p53 protein, phospho-p53 (Ser15) and p53-acetylated
lysines were captured by the coated antibody in three independent
experiences. Following extensive washing procedures, mouse anti-
bodies were added to detect the captured total p 53 protein, phospho-
p53 (Ser15) and p53-acetylated lysines, respectively. Horseradish
peroxidase substrate-linked anti-mouse antibody was then used to
recognize the bound detection antibody. Horseradish peroxidase
substrate (3,3’,5,5’-tetramethylbenzidine base) was added to develop
color. The magnitude of absorbance for this developed color (450 nm)
was proportional to the quantity of total p53 protein, phospho-p53
(Ser15) and p53-acetylated lys ines, respectively.
Statistical analyses
Descriptive results of c ontinuous variable s are expressed as mean
±
s.d.
Before statistical analysis, normal distr ibution and homogeneity of the
variances were evaluated using Levene’s test. Variables were given a base
log
10
-transformation if necessary. These parameters were analyzed on a
log scale an d tested for significan ce on that scale. The anti-log-
transformed values of the means (geometric mean) are reported in the
following table. The relation between variables was tested using
Pearson’s test and stepwise multiple linear regression analysis. ANOVA
and the unpaired t-tests were used for comparisons of quantitative
variables. The general linear model was also used to calc ulate OM at p53
gene expression after adjusting for clinical variables. The statistical
analyses were performed using the program SPSS (version 13.0; SPSS Inc.,
Chicago, IL, USA).
RESULTS
Adipose p53 and obesity-associated inflammation
Gene expression of p53, MDM2, IRS1, GLUT4, LEP, TNF-a and CD68
was quantified in the SC and OM adipose tissue from a final
cohort of 230 subjects (76 paired samples). The anthropometric
and metabolic characteristics of the participants are summarized
in Table 1.
For the whole cohort, OM p53 was positively associated with
BMI (r ¼ 0.246, P ¼ 0.001, Figure 1a and Supplementary Figure S1a
in the online appendix) and percent fat mass (r ¼ 0.199, P ¼ 0.006,
Table 2). In contrast, SC p53 was not associated with BMI
(Figure 1b). No sex-related differences for these associations were
found. On the other hand, MDM2 (Supplementary Figure S1b)
gene expression levels were downregulated in adipose
from obese subjects (Table 1), concomitantly with GLUT4
(Supplementary Figure S1c) and IRS1 (Supplementary Figure
S1d). Both OM and SC MDM2 mRNA were associated with GLUT4
(r ¼ 0.343, Po0.0001 and r ¼ 0.307, P ¼ 0.012), IRS1 (r ¼ 0.391,
Po0.0001 and r ¼ 0.403, P ¼ 0.022) and the gene expression of
fatty acid synthase in OM and SC samples of adipose tissue
(r ¼ 0.331, Po0.0001 and r ¼ 0.432, P ¼ 0.014, respectively). Inter-
estingly, OM p53 expression also correlated with MDM2 (r ¼ 0.291,
P ¼ 0.008) in non-obese individuals (but not in obese subjects),
and with TNF-a (r ¼ 0.36, P
¼ 0.001) and CD68 (r ¼ 0.246, P ¼ 0.004)
gene expressions in OM adipose tissue (Table 2).
There is evidence indicating that mRNA may not necessarily
predict the translated protein levels. In this regard, measurements
of total p53 protein ( þ 11%, P ¼ 0.04, Supplementary Figure S2a),
phosphorylated (
Ser15
P)-p53 ( þ 16%, P ¼ 0.08; Supplementary
Figure S2b) and acetylated (A)-p53 ( þ 17%, P ¼ 0.01;
Supplementary Figure S2c) in human fat depots from a
subpopulation of 28 individuals (28 paired samples of OM and
SC adipose tissue) corroborated that the p53 levels in OM (but not
in SC fat) were increased with obesity. However, the activation
state of p53 (that is, the ratio
Ser15
P-p53/p53 and A-p53/p53) in
OM and SC human adipose tissue was not significantly related to
the parameters of obesity.
SC p53 expression was inversely related to cholesterol
(r ¼0.212, P ¼ 0.029) and low-density cholesterol (r ¼0.207,
P ¼ 0.041). The same relationship was noticed between OM p53
gene expression, circulating cholesterol (r ¼0.257, P ¼ 0.02) and
low-density cholesterol (r ¼0.3, P ¼ 0.01) in the subset of obese
individuals (n ¼ 115) (Table 2). Of note, OM (but not SC) MDM2 was
also associated with cholesterol (r ¼ 0.188, P ¼ 0.035) and low-
density lipids (r ¼ 0.204, P ¼ 0.026).
p53 gene expression in mesenteric adipose tissue from mice
models
In concordance with values of p53 in humans and concomitantly
with inflammatory markers such as s100a8 ( þ 124%, P ¼ 0.016),
lbp ( þ 125%, P ¼ 0.029) and fabp4 ( þ 79.5%, P ¼ 0.044), p53
mRNA was significantly increased in mesenteric fat from diet-
induced obese mice ( þ 57.6%, P ¼ 0.003) when compared with
lean mice (Figure 1c). Accordingly, p53 gene expression was
significantly increased ( þ 29.8%, P ¼ 0.004) in fat depots from
high-fat diet-fed mice when compared with normal chow diet-fed
mice (Figure 1d).
p53 during in vitro adipogenesis and isolated fat cells. The
monitoring of p53 gene expression and protein levels during
the adipogenic maturation of human pre-adipocytes and 3T3-L1
cells to MAs disclosed the decreased expression of p53 (Figure 2a
and Supplementary Figure S3a in the online appendix, respe-
ctively). Upon differentiation, the human pre-adipocytes developed
microscopically visible lipid droplets starting on the seventh
day. Concomitantly, there was a decrease in p53 expression
( 39.3%, P ¼ 0.038) accompanied by increased anti-inflammatory
mediators such as ADIPOQ (B107-fold,
Po0.0001) but decreased
pro-inflammatory cytokines such as IL-6 ( 97.5%, P ¼ 0.001).
MDM2 expressions succinctly increased with differentiation
(B2-fold, P ¼ 0.002).
In agreement with these results, the stromal-vascular cell
fraction was mainly responsible for p53 gene expression in both
p53 in human adipose tissue
FJ Ortega et al
3
& 2013 Macmillan Publishers Limited International Journal of Obesity (2013) 1 9
SC and OM fat depots, as p53 mRNA was significantly lower
( 43.3%, Po0.0001) in ex vivo isolated MAs (Figure 2b).
Effects of LPS-MCM and Rs on p53 gene expression in fully
differentiated adipocytes
As expected, LPS-induced macrophage-conditioned medium (5%)
administration increased LEP (Figure 2e) and IL-6 (B16-fold,
Po0.0001), lowering the expression of adipogenic and/or anti-
inflammatory factors such as ADIPOQ (Figure 2d). On the other
hand, Rs administration (2 mM) reduced LEP (Figure 2e) and IL-6
( 64%, Po0.0001), and increased the expression of ADIPOQ
(Figure 2d) as well as
p(panTyr)
IRS1/
total
IRS1 ratios (data not shown)
in MAs. In agreement with previous findings showing triggered
p53 expression by inflammation, LPS-induced macrophage-con-
ditioned medium treatment increased ( þ 35%, Po0.0001)
whereas Rs significantly decreased ( 40%, P ¼ 0.005) p53
gene expression in cultured human SC adipocytes (Figure 2c).
LPS by itself was able to induce the upregulation of p53 gene
expression also in mature 3T3-L1 adipocytes ( þ 93%, P ¼ 0.019;
Supplementary Figure S3b).
Adipose tissue p53 and altered glucose tolerance
OM (but not SC) p53 expression levels correlated with fasting
glucose (Figure 3a), with glucose concentrations at 0
0
(r ¼0.39,
P ¼ 0.036), 90
0
(r ¼0.4, P ¼ 0.035) and 120
0
(r ¼0.45,
P ¼ 0.016, n ¼ 28) during the oral glucose tolerance test, and
was significantly associated with the area under the curve for
glucose (r ¼0.39, P ¼ 0.039) in the random subsample of 28
subjects. Interestingly, OM p53 was strongly associated with
glycated hemoglobin (r ¼0.337, Po0.0001; Figure3b), a well
known measure of integrated glucose levels. Indeed, multiple
linear regression analyses to predict OM p53 levels revealed that
glycated hemoglobin (Po0.0001) and BMI (P ¼ 0.048) contributed
independently to explain 13.7% (Po0.0001) of OM p53 variance
(Table 2) after controlling for age. This association was of
special relevance among the non-obese subjects, and was not
Table 1. Anthropometrical and biochemical characteristics of study subjects
Whole cohort Non-obese without IGT Non-obese and IGT Obese without IGT Obese and IGT P (ANOVA)
N 95 21 61 53
Sex (% women) 61.7 71.4 83.3 77.4
Age (years) 49
±
14 56
±
10 44
±
12 47
±
12 0.002
BMI (kg m
2
) 25.0
±
3.1 27.5
±
1.8 42.6
±
6.7 43.0
±
5.7 o0.0001
% fat 31.7
±
7.1 37.5
±
6.6 54.1
±
8.7 54.6
±
9.2 o0.0001
SBP (mm Hg) 125.0
±
19.5 136.7
±
17.1 134.9
±
17.9 141.6
±
18.8 o0.0001
DBP (mm Hg) 77.7
±
13.0 78.7
±
11.3 78.8
±
10.5 81.8
±
11.1 0.421
Fasting glucose (mg dl
1
) 85.7
±
10.3 135.8
±
54.2 89.1
±
10.4 131.5
±
50.3 o0.0001
Insulin (mUml
1
) 11.2
±
6.8 15.0
±
7.7 13.7
±
9.1 19.2
±
13.5 0.005
HOMA-IR 2.39
±
1.54 4.72
±
2.24 2.97
±
1.87 5.70
±
4.09 o0.0001
HbA1c (%) 5.4
±
0.5 6.9
±
2.2 5.0
±
0.6 5.8
±
1.4 o0.0001
Total cholesterol (mg dl
1
) 202.9
±
38.3 209.2
±
44.7 188.6
±
28.8 197.2
±
37.5 0.072
HDL-cholesterol (mg dl
1
) 57.0
±
16.2 55.2
±
23.6 54.4
±
13.9 60.2
±
47.5 0.743
LDL-cholesterol (mg dl
1
) 123.3
±
30.2 120.4
±
43.3 114.8
±
28.2 114.6
±
36.2 0.367
Fasting triglycerides (mg dl
1
) 104.9
±
47.0 179.3
±
100.8 106.0
±
50.3 146.3
±
86.2 o0.0001
OM p53 gene expression (AU) 0.043
±
0.017 0.035
±
0.016 0.053
±
0.022 0.050
±
0.017 0.001
SC p53 gene expression (AU) 0.040
±
0.010 0.044
±
0.012 0.039
±
0.010 0.041
±
0.013 0.585
OM MDM2 gene expression (AU) 0.039
±
0.020 0.031
±
0.010 0.033
±
0.020 0.029
±
0.016 0.034
SC MDM2 gene expression (AU) 0.033
±
0.009 0.038
±
0.015 0.022
±
0.006 0.026
±
0.006 o0.0001
OM GLUT4 gene expression (AU) 0.043
±
0.035 0.020
±
0.018 0.024
±
0.022 0.017
±
0.010 o0.0001
SC GLUT4 gene expression (AU) 0.060
±
0.034 0.042
±
0.014 0.046
±
0.025 0.030
±
0.023 o0.0001
OM IRS1 gene expression (AU) 0.018
±
0.012 0.012
±
0.007 0.011
±
0.004 0.010
±
0.006 o0.0001
SC IRS1 gene expression (AU) 0.015
±
0.010 0.015
±
0.006 0.012
±
0.007 0.009
±
0.003 0.049
OM LEP gene expression (AU) 0.164
±
0.164 0.268
±
0.187 0.276
±
0.130 0.327
±
0.137 o0.0001
SC LEP gene expression (AU) 0.622
±
0.311 0.757
±
0.236 0.954
±
0.309 0.856
±
0.340 0.001
OM TNF-a gene expression (AU) 0.004
±
0.003 0.002
±
0.002 0.005
±
0.003 0.005
±
0.002 0.05
SC TNF-a gene expression (AU) 0.005
±
0.004 0.003
±
0.001 0.003
±
0.001 0.004
±
0.002 0.017
OM CD68 gene expression (AU) 0.163
±
0.102 0.146
±
0.096 0.179
±
0.076 0.23
±
0.113 0.011
SC CD68 gene expression (AU) 0.241
±
0.105 0.201
±
0.109 0.186
±
0.095 0.226
±
0.115 0.275
Subpopulation (A) Non-obese without IGT (B) Obese without IGT (C) Obese and IGT Student t (A) vs (B)
N 8911
OM p53 (OD) 0.099
±
0.006 0.110
±
0.013 0.105
±
0.015 0.044
SC p53 (OD) 0.110
±
0.008 0.112
±
0.008 0.109
±
0.008 0.572
OM P-p53 (OD) 0.106
±
0.015 0.123
±
0.021 0.110
±
0.017 0.078
SC P-p53 (OD) 0.111
±
0.012 0.121
±
0.018 0.114
±
0.005 0.248
OM A-p53 (OD) 0.374
±
0.035 0.439
±
0.051 0.420
±
0.030 0.009
SC A-p53 (OD) 0.415
±
0.074 0.476
±
0.074 0.483
±
0.081 0.156
Abbreviations: AU, arbitrary units; BMI, body mass index; CD68, cluster differentiation 68 molecule; DBP, diastolic blood pressure; GLUT4, glucose transporter
type 4; HbA1c, glycated hemoglobin; HDL, high-density lipoprotein; HOMA-IR, homeostasis model assessment of insulin resistance value, calculatedinall
subjects as (glucose (mmol l
1
) insulin (mU l
1
)/22.5); IGT, impaired glucose tolerance; IRS1, insulin receptor soluble 1; LDL, low-density lipoprotein;
LEP, leptin; MDM2, mouse double minute 2; OD, optical density; OM, omental fat; p53, tumor protein 53; SBP, systolic blood pressure; SC, subcutaneous fat;
TNF-a, tumor necrosis factor-alpha. Bold refers to significant differences between groups. ANOVAs P-values depict significant variations among all groups.
Data are means
±
s.d. Non-obese, BMIo30 kg m
2
; obese, BMIX30 kg m
2
.
p53 in human adipose tissue
FJ Ortega et al
4
International Journal of Obesity (2013) 1 9 & 2013 Macmillan Publishers Limited
0.12
r=0.246
p=0.001
r=-0.05
p=0.596
0.10
SC p53 gene expression (AU)
OM p53 gene expression (AU)
0.08
0.06
0.04
0.02
0.00
20 30 40
Body Mass Index (Kg m
–2
)
Lean mice Obese mice NC-feed mice
#
*
5.0
4.0
3.0
3.0
2.0
2.0
Mesenteric p53 gene
expression (AU)
Mesenteric p53 gene
expression (AU)
1.0
1.0
1.5
2.5
0.0
0.0
0.5
HF-feed mice
50 60 20 30 40
Body Mass Index (Kg m
–2
)
50 60
0.12
0.10
0.08
0.06
0.04
0.02
0.00
Figure 1. Upper panels: Linear relationships between body mass index (kg m
2
) and omental (OM; a) and subcutaneous (SC; b) tumor protein
53 (p53) gene expression levels. Values for men are represented as empty circles (J), and women by empty diamonds (B). Lower panels:
Mean
±
2.0 s.e. for the mean of p53 gene expression in high-fat diet-induced obese and lean mice (c). The acute effect of high-fat diet on p53
gene expression levels in adipose tissue was also analyzed (d). *Po0.05 and
#
Po0.0001 for comparisons with the control group.
Table 2. Correlations and multiple linear regressions for tumor p53 gene expressions in OM fat depots and study variables
All subjects Non-obese (BMIo30 kg m
2
) Obesity (BMIX30 kg m
2
)
Correlations rP r P r P
Age (years) 0.137 0.058 0.109 0.276 0.092 0.39
BMI (kg m
2
) 0.246 0.001 0.08 0.425 0.016 0.881
Fat mass (%) 0.199 0.006 0.012 0.907 0.036 0.738
Fasting glucose (mg dl
1
) 0.111 0.151 0.244 0.017 0.084 0.478
HbA
1c
(%) 0.337 o0.0001 0.381 o0.0001 0.238 0.082
Total cholesterol (mg dl
1
) 0.094 0.214 0.107 0.3 0.257 0.02
HDL-cholesterol (mg dl
1
) 0.073 0.343 0.21 0.042 0.049 0.673
LDL-cholesterol (mg dl
1
) 0.139 0.072 0.072 0.493 0.3 0.01
Fasting triglycerides (mg dl
1
) 0.056 0.463 0.161 0.12 0.01 0.932
OM MDM2 gene expression (AU) 0.069 0.422 0.291 0.008 0.096 0.489
OM GLUT4 gene expression (AU) 0.144 0.093 0.289 0.008 0.292 0.03
OM IRS1 gene expression (AU) 0.2 0.017 0.339 0.003 0.27 0.026
OM LEP gene expression (AU) 0.159 0.072 0.034 0.754 0.025 0.882
OM TNF-a gene expression (AU) 0.36 0.001 0.508 0.002 0.286 0.07
OM CD68 gene expression (AU) 0.246 0.004 0.356 0.001 0.016 0.909
SC p53 gene expression (AU) 0.18 0.119 0.223 0.22 0.168 0.275
Multiple linear regression Beta P
Age (years) 0.022 0.798
BMI (kg m
2
) 0.157 0.048
HbA
1c
(%) 0.317 o0.0001
Adjusted R
2
13.7 (
P
o0.0001)
Abbreviations: AU, arbitrary units; BMI, body mass index; CD68, cluster differentiation 68 molec ule; GLUT4, glucose transporter type 4; HbA
1c
, glycated
hemoglobin; IRS1, insulin receptor soluble 1; LEP, leptin; MDM2, mouse double minute 2; OM, omental fat; p53, protein 53; SC, subcutaneous fat; TNF-a,
tumor necrosis factor-alpha. Beta is the standardized regression coefficient that allows evaluating the relative significance of each independent variable
in multiple linear regression analysis. Adjusted R
2
expresses the percentage of the variance explained by the independent variables in the different models
(i.e., 0.50 is 50%). Significant associations are indicated in bold.
p53 in human adipose tissue
FJ Ortega et al
5
& 2013 Macmillan Publishers Limited International Journal of Obesity (2013) 1 9
age-dependent (Table 2). Concurrently, p53 was positively
associated with the gene expression of insulin action biomarkers
in OM adipose tissue, such as GLUT4 and IRS1, especially among
non-obese participants (r ¼ 0.289, P ¼ 0.008 and r ¼ 0.339,
P ¼ 0.003, respectively; Table 2).
p53 gene expression in mesenteric adipose tissue from mice
models
Of note, GLP-1R KO mice, a genetic mouse model of glucose
intolerance,
25
showed decreased expressions of mesenteric p53
( 45.4%, P ¼ 0.017) when compared with wild-type mice
(Figure 3c). This was in agreement with the notion that impaired
glucose tolerance is associated with decreased p53 expression in
human OM adipose tissue.
Effects of surgery-induced weight loss on SC p53 gene expression
The characteristics of the subjects included in this longitudinal
study are shown in Supplementary Table S1. In an independent
cohort of morbidly obese women (n ¼ 7), bariatric surgery-
induced weight loss (60.1
±
11.7 vs 49.5
±
3.4 kg m
2
) led to
significantly increased SC p53 ( þ 44.9%, P ¼ 0.029; Supplementary
Figure S4a) gene expression levels despite the surgery-induced
downregulation of biomarkers of inflammation (I L-6, 89%,
P ¼ 0.036) and macrophage infiltration (CD14, 54.2%, P ¼ 0.02)
(Supplemental Table S1). On the other hand, i ncreased GLUT4
( þ 89.9%, P ¼ 0.01; Supplementary Figure S4c) and IRS1
( þ 99.1%, P ¼ 0.007; Supplementary Figure S4d) as well as
decreased LEP ( 55.3%, P ¼ 0.024; Supplementary Figure S4e)
expressions were concomitant to decreased weight and
increased insulin sensitivity (hyperinsulinemic euglycemic M value).
No significant changes for MDM2 gene expression were found
(Supplementary Figure S4b).
Effects of metformin on SC p53 gene expression
In the context of its well known enhanced AMP-activated protein
kinase effects, metformin (10 mmol l
1
) led to significantly
0.08
0.07
0.06
p53 gene
expression (AU)
p53 gene
expression (AU)
0.05
0.04
0.03
0.02
0.01
0.00
0.07
0.06
0.05
0.04
0.03
0.02
0.01
0.00
SVCs
*
Mature
adipocytes
#
Pre-
adipocytes
Adipocytes
(7th day)
Mature adipocytes
(14th day)
Control
0.0
2.0
4.0
6.0
8.0
10.0
*
#
+MCM (5%)
#
+Rs (2 uM)
Control
+MCM (5%)
+Rs (2 uM)
Control
+MCM (5%)
+Rs (2 uM)
*
*
*
0.0
0.5
p53 gene
expression (AU)
AdipoQ gene
expression (AU)
LEP gene
expression (AU)
1.0
1.5
4.0
3.0
2.0
1.0
0.0
Figure 2. Mean
±
2.0 s.e. for the mean of p53 gene expression at the 0th, 7th and 14th days after inducing adipogenesis (a), and in the cellular
fractions of adipose tissue (stromal-vascular cells and MAs) isolated from fresh human adipose tissue ( b). Lower panels represent the effect in
p53 (c), ADIPOQ (d) and LEP (e) gene expression of LPS-induced macrophage-conditioned medium and Rs in MAs. *Po0.05 and
#
Po0.0001 for
comparisons with pre-adipocytes/non-treated cells.
0.12
r=-0.111
p=0.151
r=-0.337
p <0.0001
0.10
0.08
OM p53 gene
expression (AU)
0.06
0.04
0.02
0.00
0.12
0.10
0.08
OM p53 gene
expression (AU)
0.06
0.04
0.02
0.00
60 80 100 120
140
Fasting glucose (mg dL
–1
) Hb1Ac (%)
NC-fed
WT mice
NC-fed
GLP-1R KO mice
456789
10
3.0
2.0
Mesenteric p53 gene
expression (AU)
1.0
*
1.5
2.5
0.0
0.5
Figure 3. Linear relationships between fasting glucose (a) and glycated hemoglobin (Hb1Ac, b) with p53 gene expression levels. Values for
men are represented as empty circles (J); and women as empty diamonds (B). Striped bars (c) represent mean
±
2.0 s.e. for the mean of p53
gene expression in mesenteric (Mes) adipose tissue from normal chow-fed GLP-1R KO transgenic and wild-type mice. *Po0.05 for
comparisons between group.
p53 in human adipose tissue
FJ Ortega et al
6
International Journal of Obesity (2013) 1 9 & 2013 Macmillan Publishers Limited
increased p53 gene expression levels ( þ 42%, P ¼ 0.018;
Supplementary Figure S5) ex vivo, in explants of human SC
adipose tissue.
Effects of glucose on p53 gene expression in 3T3-L1.Togain
insight into the potential underlyi ng molecular links between
increased glucose and p53 gene expressions in adipocytes, the
present study performed further experiments in 3T3-L1 pre-
adipocytes cultured with different concentrations of glucose
(low: 10; normal: 20 and high: 200 m
M). Interestingly, high
glucose concentrations significantly d ownregulated p53
( 27%, Po0.0001; Figure 4a), in parallel to the expression
pattern of biomarkers for insulin sensit ivity such as IRS1
( 44%, Po0.0001; Figure 4c) and AdipoQ ( 97%, P ¼ 0.005;
Figure 4e ).
DISCUSSION
The present work aimed at evaluating p53 in human adipose
tissue in association with obesity-associated inflammation
and insulin resistance. We found a dual relationship: the
upregulation of p53 linked to inflammation seemed to be
counterbalanced by the high glucose and insulin resistance-
dependent downregulation.
Inflammation-dependent effects on adipose tissue p53 gene
expression
Inflammation is well known to upregulate p53. Reactive oxygen
species (ROS)-induced p53 activation causes the NF-kB-depen-
dent induction of inflammatory cyt okines and cell d eath.
26
As
NF-kB induction occurs as a response to stress and p53 arrests
cells in G1/S (where DNA repair may be initiated ), the induction
of p53 by NF-kB could be a mechanism by which cells can
recover from stressors.
27
The upregulation of p53 in the OM
(but not in SC) adipose tissue with obesity, with its well known
chronic low-grade inflammatory activity,
4,6
could be interpreted
in this conte xt as further indicated by positive associations with
markers of inflammation (TNF-a) and macrophage infiltration
(CD68) in the OM fat depots. Then, the findings in humans
confirmed increased p53 expression in the adipose tissue of Ay
mice, a model in which increased inflammatory activity is
prominent. We here also r eplicated increased adipose tissue
p53 in high fat-fed mice. Thus, in obese adipose tissue and high
fat-fed mice increased inflammatory cytokines may account for
enhanced p53, whereas in differ entiated adipocytes increased
anti-inflammatory adipokines like adiponectin and the decreased
inflammation (which is necessary for adipogenesis) may explain the
in vitro p53 downregulation during differentiation.
Adipose tissue senescence was recently demonstrated in obese
subjects by the presence of shortened telomeres.
7
This may
increase the local production of pro-inflammatory molecules and
the upregulation of p53 in adipose tissue. In line with these
findings, p53 expression was higher in the most inflammatory
cellular counterpart of the adipose tissue: the stromal-vascular
fraction. Furthermore, LPS-macrophage-conditioned medium led
to raised p53 expression in MAs, as expected because cytokines
are known to activate p53 expression in concert with NF-kB.
28
On
the other hand, Rs, with its well known anti-inflammatory activity,
led to the downregulation of p53 in MAs (current findings).
Of note, MDM2 expression levels in adipose w ere positively
associated with the gene markers of lipogenesis (that is, FASN
and ADIPOQ) and insulin sensitivity (that is, IRS1 and GLUT4),
decreasing concomitantly with obesity. The MDM2 protein is
one of the most important p53-ubiquitin lig ases that induces
p53 polyubiquitination and degradation.
29
Although increased
p53 levels led to p 53-mediated cell death, MDM2 is requir ed t o
maintain p53 at low levels and to suppress i ts ability to induce
cell death in proliferating and qui escent cell s.
30
The opposite
expression pattern of MDM2 and p53 in OM fat depots, as
well as during adipogenesis, highlights its obesity-related
modulation.
High glucose and insulin resistance effects on adipose tissue p53
expression
p53 expression in OM adipose tissue was inversely associated with
glycated hemoglobin, a measure of integrated glucose levels. p53
gene expressions were also negatively linked to insulin resistance
and positively with the expression of insulin sensitivity-related
genes (GLUT4 and IRS1). Moreover, the gastric bypass bariatric
surgery, known as an effective approach for achieving weight
loss in obese patients,
31
improving and even completely solving
most obesity-associated complications, particularly insulin resistance
and type 2 diabetes 20, increased SC p53 expression, concomitantly
with increased insulin sensitivity and decreased inflammation (IL-6)
and macrophage infiltration (CD14). These findings were confirmed
in an animal model of glucose intolerance, the GLP-1R KO mice 25,
0.010
0.008
#
#
*
*
0.003
0.002
0.001
0.000
*
#
0.0E0
2.0E-5
1.5E-5
1.0E-5
5.0E-6
0.0E0
0.0000
adipoq gene
expression (AU)
0.0005
0.0010
0.0015
0.0020
irs1 gene
expression (AU)
glut4 gene
expression (AU)
Low
glucose
Normal
glucose
High
glucose
Low
glucose
Normal
glucose
High
glucose
Low
glucose
Normal
glucose
High
glucose
Low
glucose
Normal
glucose
High
glucose
Low
glucose
Normal
glucose
High
glucose
*
#
1.0E-4
2.0E-4
3.0E-4
4.0E-4
5.0E-4
il6 gene
expression (AU)
p53 gene
expression (AU)
6.0E-4
7.0E-4
0.006
0.004
0.002
0.000
Figure 4. Mean
±
2.0 s.e. for the mean of p53 (a), Il-6 (b), IRS1 (c), GLUT4 (d) and ADIPOQ (e) gene expression levels in 3T3-L1 pre-adipocytes
cultured under conditions of ‘low (10 m
M), ‘normal’ (20 mM) and ‘high’ glucose (200 mM). *Po0.05 and
#
Po0.0001 for comparisons with cells
cultured with ‘normal’ glucose concentration in the medium.
p53 in human adipose tissue
FJ Ortega et al
7
& 2013 Macmillan Publishers Limited International Journal of Obesity (2013) 1 9
that also showed decreased expressions of p53 in adipose tissue.
Furthermore, the relatively high glucose levels of in vitro adipocyte
differentiation led to decreased p53 levels despite increased IL-6
gene expression ( þ 200%, Po0.0001; Figure 4b). Thus, according
to our results the higher the glucose concentration, the lower the
p53 gene expression in adipocytes.
On the other hand, AMP-activated protein kinase constitutes a
major metabolic regulator of energy homeostasis that upregulates
p53, ensuring the arrest of the cel l c ycle in peri ods of glucose
deprivatio n.
32
Interestingly, metformin, known to activate
AMP-activated protein kinase, led to raised p53 gene
expression in adipose tissue. In line with t his effect,
metformin also led to p53 induction in parallel to Bcl-2
downregulation in melanoma cells, leading to cell cycle
arrest and apoptosis.
33
All these results suggest that hyperglycemia and/or insulin
resistance are associat ed with decreased apopt osis. Previous
studies have shown that hyperglycemia inhibited rat vascular
smooth muscle cell apoptosis in vi tro
34
and in patients with
diabetes.
35
When nondiabetic vascular smooth muscle cells
were preincubated with increasing concentrations of glucose
for 48 h, the y became resist ant to apoptosis, similar to the
diabetic cells.
35
The decrease in apoptosis in vascular smooth
muscle cells was re flected in an increasing artery-media
thicknessinpatientswithdiabetes.Thedecreasedapoptosis
with worsening of hyperglycemi a here found (at least at the
level o f adipose tissue p53 gene expression) might have clinical
implications. Type 2 diabetic patients are known to lose less
weight after dieting than nondiabe tic obe se sub jects.
36,37
Here
we postulate that the decreased p53 expression induced by
hyperglycemia might limit adipose tissue renewal and
resistance to weight loss.
In summary, factors linked to inflammation, such as weight gain,
aging and high-fat diet, among others, amplify p53 gene
expression, whereas hyperglycemia leads to decreased p53
expression in human adipose tissue. The specific involvement of
p53 mechanisms in these processes in human adipose tissue
should be investigated further.
CONFLICT OF INTEREST
The authors declare no conflict of interest.
ACKNOWLEDGEMENTS
We grea tly appreciate th e technical ass istan ce of Gerard Pardo, Ester Guerra,
Oscar Rovira and Roser Rodriguez (Unit of Diabetes, Endocrinology and Nutrition,
Institut d’Investigacio
´
Biome
`
dica de Girona; Hospital Universitari de Girona
Dr Jos ep Trueta). The work of all the members of the multidisciplinary obesity
team of the Clı
´
nica Universitaria de Navarra is also gratefully acknowledged. This
work was supported by research grants from the Ministerio de Educa cio
´
ny
Ciencia (SAF2011-00214) and the Instituto de Salud Carlos III (ISCIIIRETIC RD06,
CIBERObN). RB is a recipient of funding from the Agence Nationale de la
Recherche (Florinfl am and Flora dip programs ) as well as f rom L’Instit ute Nationa le
du Di abe
`
te.
AUTHOR CONTRIBUTIONS
All authors of this manuscript have directly participated in the execution and analysis
of the study and have approved the final version submitted. FJO designed the study,
analyzed the biochemical variables, performed the statistical analysis and wrote the
manuscript. JMM-N, DM and MS analyzed the biochemical variables. JIR-H obtained
the biopsies, the anthropometrical characteristics and the written consent of
volunteers. EL and RB were responsible for mice, and the mouse models of obesity
and high-fat diet. GM obtained subcutaneous fat samples from morbid obese
volunteers before and after bariatric surgery. WR, RB, FT, GF and GM provided
important intellectual content. AZ also carried out the conception of the study. JMF-R
carried out the conception and coordination of the study, and contributed to writing
the manuscript.
DISCLAIMER
The contents of this m anuscript have not been copyrighted or published
previously. There are no directly related manuscripts or abstracts, published or
unpublished, by on e or more a uthors of this manuscript. The contents of this
manuscript are not now under consideration for p ublication elsewhere. The
submitted manuscript nor any similar manuscript, in whole or in part, will n ot be
copyrighted, submi tted or published elsewhere while the Journal is under
consideration.
REFERENCES
1 Levine AJ, Oren M. The first 30 years of p53: growing ever more complex. Nat Rev
Cancer 2009; 9: 749–758.
2 Benoit V, de Moraes E, Dar NA, Taranchon E, Bours V, Hautefeuille A et al.
Transcriptional activation of cyclooxygenase-2 by tumor suppressor p53 requires
nuclear factor-kappaB. Oncogene 2006; 25: 5708–5718.
3 Maddocks OD, Vousden KH. Metabolic regulation by p53. J Mol Med (Berl) 2011;
89: 237–245.
4 Gregor MF, Hotamisligil GS. Inflammatory mechanisms in obesity. Annu Rev
Immunol 2011; 29: 415–445.
5 Shimizu I, Yoshida Y, Katsuno T, Minamino T. Adipose tissue inflammation in
diabetes and heart failure. Microbes Infect 2013; 15: 11–17.
6 Wellen KE, Hotamisligil GS. Obesity-induced inflammatory changes in adipose
tissue. J Clin Invest 2003; 112: 1785–1788.
7 Moreno-Navarrete JM, Ortega F, Sabater M, Ricart W, Fernandez-Real JM.
Telomere length of subcutaneous adipose tissue cells is shorter in obese and
formerly obese subjects. Int J Obes (Lond) 2010; 34: 1345–1348.
8 Minamino T, Orimo M, Shimizu I, Kunieda T, Yokoyama M, Ito T et al. A crucial role
for adipose tissue p53 in the regulation of insulin resistance. Nat Med 2009; 15:
1082–1087.
9 Shaulsky G, Goldfing er N, Peled A, Rotter V. Involvement of wild-type p53 in
pre-B-cell differentiation in vitro. Proc Natl Acad Sci USA 1991; 88: 8982–8986.
10 Wang X, Kua HY, Hu Y, Guo K, Zeng Q, Wu Q et al. p53 functions as a negative
regulator of osteoblastogenesis, osteoblast-dependent osteoclastogenesis, and
bone remodeling. J Cell Biol 2006; 172: 115–125.
11 Zambetti GP, Horwitz EM, Schipani E. Skeletons in the p53 tumor suppressor
closet: genetic evidence that p53 blocks bone differentiation and development.
J Cell Biol 2006; 172: 795–797.
12 Bazuine M, Stenkula KG, Cam M, Arroyo M, Cushman SW. Guardian of corpulence:
a hypothesis on p53 signaling in the fat cell. Clin Lipidol 2009; 4: 231–243.
13 Kichina J, Green A, Rauth S. Tumor suppressor p53 down-regulates tissue-specific
expression of tyrosinase gene in human melanoma cell lines. Pigment Cell Res
1996; 9: 85–91.
14 Schwartzenberg-Bar-Yoseph F, Armoni M, Karnieli E. The tumor suppressor p53
down-regulates glucose transporters GLUT1 and GLUT4 gene expression. Cancer
Res 2004; 64: 2627–2633.
15 Molchadsky A, Shats I, Goldfinger N, Pevsner-Fischer M, Olson M, Rinon A et al.
p53 plays a role in mesenchymal differentiation programs, in a cell fate
dependent manner. PLoS One 2008; 3: e3707.
16 de la Torre AJ, Rogoff D, White PC. p53 and cellular glucose uptake. Endocr Res
2013; 38: 32–39.
17 Budanov AV, Karin M. p53 target genes sestrin1 and sestrin2 connect genotoxic
stress and mTOR signaling. Cell 2008; 134: 451–460.
18 Bensaad K, Vousden KH. Savior and slayer: the two faces of p53. Nat Med 2005; 11:
1278–1279.
19 Bunnell BA, Flaat M, Gagliardi C, Patel B, Ripoll C. Adipose-derived stem cells:
isolation, expansion and differentiation. Methods 2008; 45: 115–120.
20 Mingrone G, Panunzi S, De Gaetano A, Guidone C, Iaconelli A, Leccesi L et al.
Bariatric surgery versus conventional medical therapy for type 2 diabetes. N Engl J
Med 2013; 366: 1577–1585.
21 Moreno-Navarrete JM, Ortega FJ, Rodriguez-Hermosa JI, Sabater M, Pardo G,
Ricart W et al. OCT1 expression in adipocytes could contribute to increased
metformin action in obese subjects. Diabetes 2011; 60: 168–176.
22 Cani PD, Amar J, Iglesias MA, Poggi M, Knauf C, Bastelica D et al. Metabolic
endotoxemia initiates obesity and insulin resistance. Diabetes 2007; 56:
1761–1772.
23 Burcelin R, Crivelli V, Dacosta A, Roy-Tirelli A, Thorens B. Heterogeneous metabolic
adaptation of C57BL/6J mice to high -fat diet. Am J Physiol Endocrinol Metab 2002;
282: E834–E84 2.
24 Ortega FJ, Vazquez-Martin A, Moreno-Navarrete JM, Bassols J, Rodriguez-Hermosa
J, Girones J et al. Thyroid hormone responsive Spot 14 increases during
differentiation of human adipocytes and its expression is down-regulated in
obese subjects. Int J Obes (Lond) 2010; 34: 487–499.
p53 in human adipose tissue
FJ Ortega et al
8
International Journal of Obesity (2013) 1 9 & 2013 Macmillan Publishers Limited
25 Scrocchi LA, Brown TJ, MaClusky N, Brubaker PL, Auerbach AB, Joyner AL et al .
Glucose intolerance but normal satiety in mice with a null mutation in the
glucagon-like peptide 1 receptor gene. Nat Med 1996; 2: 1254–1258.
26 Ryan KM, Ernst MK, Rice NR, Vousden KH. Role of NF-kappaB in p53-mediated
programmed cell death. Nature 2000; 404: 892–897.
27 Wu H, Lozano G. NF-kappa B activation of p53.A potential mechanism for
suppressing cell growth in response to stress. J Biol Chem 1994; 269:
20067–20074.
28 Webster GA, Perkins ND. Transcriptional cross talk between NF-kappaB and p53.
Mol Cell Biol 1999; 19: 3485–3495.
29 Jackson MW, Berberich SJ. Constitutive mdmx expression during cell growth,
differentiation, and DNA damage. DNA Cell Biol 1999; 18: 693–700.
30 Marine JC, Francoz S, Maetens M, Wahl G, Toledo F, Lozano G. Keeping p53 in
check: essential and synergistic functions of Mdm2 and Mdm4. Cell Death Differ
2006; 13: 927–934.
31 Chang J, Wittert G. Effects of bariatric surgery on morbidity and mortality in
severe obesity. Int J Evid Based Healthc 2009; 7: 43–48.
32 Thoreen CC, Sabatini DM. AMPK and p53 help cells through lean times. Cell Metab
2005; 1: 287–288.
33 Janjetovic K, Harhaji-Trajkovic L, Misirkic-Marjanovic M, Vucicevic L, Stevanovic D,
Zogovic N et al. In vitro and in vivo anti-melanoma action of metformin. Eur J
Pharmacol 2011; 668: 373–382.
34 Hall JL, Matter CM, Wang X, Gibbons GH. Hyperglycemia inhibits vascular smooth
muscle cell apoptosis through a protein kinase C-dependent pathway. Circ Res
2000; 87: 574–580.
35 Ruiz E, Gordillo-Moscoso A, Padilla E, Redondo S, Rodriguez E, Reguillo F et al .
Human vascular smooth muscle cells from diabetic patients are resistant
to induced apoptosis due to high Bcl-2 expres sion. Diabetes 2006; 55:
1243–1251.
36 Wing RR, Marcus MD, Epstein LH, Salata R. Type II diabetic subjects lose less weight
than their overweight nondiabetic spouses. Diabetes Care 1987; 10: 563–566.
37 Henry RR, Wallace P, Olefsky JM. Effects of weight loss on mechanisms of
hyperglycemia in obese non-insulin-dependent diabetes mellitus. Diabetes 1986; 35:
990–998.
Supplementary Information accompanies this paper on International Journal of Obesity website (http://www.nature.com/ijo)
p53 in human adipose tissue
FJ Ortega et al
9
& 2013 Macmillan Publishers Limited International Journal of Obesity (2013) 1 9
... Yahagi et al. [6] initially reported that the mRNA and protein levels of p53 were increased in the epididymal fat tissue of young, 12-weekold, genetically obese ob/ob mice. Later, several groups reported that senescence-related markers were increased in the white adipose tissue of genetically obese Ay mice, high-fat diet-induced obese mice and rats from young, 8 weeks old, to middle age, 10 months old [7][8][9][10][11]. ...
... The gene expression and protein level of p53 and p21 and SA-β-Gal staining were increased in young obese Ay mice aged 20 weeks [7]. Similar results using an obese animal model of ob/ob or diet-induced obese mice and rats, aged 8 weeks to 10 months, were reported in several articles [8][9][10][11]25]. Even human adipose tissue showed an increased p53 expression in obese subjects [9]. ...
... Similar results using an obese animal model of ob/ob or diet-induced obese mice and rats, aged 8 weeks to 10 months, were reported in several articles [8][9][10][11]25]. Even human adipose tissue showed an increased p53 expression in obese subjects [9]. ...
Article
Full-text available
Obesity has become a global medical problem. The upregulation of senescence-related markers in adipose tissue may cause impairment of adipose tissue and disorders of systemic metabolism. Weight control through diet has been found to ameliorate senescence in the adipose tissue. Exercise is also important in maintaining a healthy lifestyle, however, very few researchers have examined the relationship between senescence-related markers in adipose tissue. Dietary restriction is also reported to have a legacy effect, wherein the effects are maintained for some periods after the termination of the intervention. However, very few researchers have examined the relationship between exercise and senescence-related markers in adipose tissue. Besides, there is no study on the long-term effects of exercise. Hence, we investigated whether the exercise could change the expression of senescence-related genes in the visceral adipose tissue of young mice and whether there was a legacy effect of exercise for 10 weeks after the termination of exercise. Four-week-old male ICR mice were assigned to one of the three groups: 20 weeks of sedentary condition, 20 weeks of voluntary wheel running exercise, or 10 weeks of exercise followed by 10 weeks of sedentary condition. The mice showed decreased expression in genes related to senescence and senescence-associated secretory phenotype, such as p53, p16, and IL-6, in the visceral adipose tissue in response to exercise. These effects were maintained for 10 weeks after the mice stopped exercising. Our study is the first report that exercise reduces the expression of senescence-related genes in the visceral adipose tissue of young mice, and that exercise causes the legacy effect.
... This, and luciferase reporter assays on the Srebf1 and fatty acid synthase (Fasn) promoter, led the authors to conclude that p53 represses the lipogenic Srebf1 pathway in adipocytes [65]. Following this initial finding, p53 mRNA and/or protein induction in visceral WAT of obese mice [66][67][68][69], rats [70], and humans [71,72] was observed. While its induction in obesity was reproduced in many studies, the mechanisms of p53 activation and its role on adipocyte function remain inconclusive, although a significant amount of research has been done in this field. ...
... Although upregulation of both p53 activity and inflammation is commonly observed in WAT of obese rodents and humans, the causal relationship remains unclear. A positive correlation between p53 expression, inflammation (e.g., TNF gene expression and CD68-positive macrophage infiltration) and body mass index (BMI) was observed in the omental (and not in subcutaneous) WAT depot in a study involving a cohort of 230 human subjects [72]. In support of inflammation-induced p53 activity postulated by the authors, in vitro experiments showed that the anti-inflammatory effects of endogenously produced adiponectin or of rosiglitazone treatment reduced p53 expression in differentiated adipocytes. ...
... In support of inflammation-induced p53 activity postulated by the authors, in vitro experiments showed that the anti-inflammatory effects of endogenously produced adiponectin or of rosiglitazone treatment reduced p53 expression in differentiated adipocytes. In line, addition of LPS-treated macrophage-conditioned medium increased p53 expression in differentiated adipocytes [72]. Importantly, Minamino et al. also analyzed visceral WAT from nondiabetic and diabetic human subjects and showed increased senescence, p53 induction, and increased TNF and CCL2 expression in diabetic patients [71]. ...
Article
Full-text available
As a tumor suppressor and the most frequently mutated gene in cancer, p53 is among the best-described molecules in medical research. As cancer is in most cases an age-related disease, it seems paradoxical that p53 is so strongly conserved from early multicellular organisms to humans. A function not directly related to tumor suppression, such as the regulation of metabolism in nontransformed cells, could explain this selective pressure. While this role of p53 in cellular metabolism is gradually emerging, it is imperative to dissect the tissue-and cell-specific actions of p53 and its downstream signaling pathways. In this review, we focus on studies reporting p53's impact on adipocyte development, function, and maintenance, as well as the causes and consequences of altered p53 levels in white and brown adipose tissue (AT) with respect to systemic energy homeostasis. While whole body p53 knockout mice gain less weight and fat mass under a high-fat diet owing to increased energy expenditure, modifying p53 expression specifically in adipocytes yields more refined insights: (1) p53 is a negative regulator of in vitro adipogenesis; (2) p53 levels in white AT are increased in diet-induced and genetic obesity mouse models and in obese humans; (3) functionally, elevated p53 in white AT increases senescence and chronic inflammation, aggravating systemic insulin resistance; (4) p53 is not required for normal development of brown AT; and (5) when p53 is activated in brown AT in mice fed a high-fat diet, it increases brown AT temperature and brown AT marker gene expression, thereby contributing to reduced fat mass accumulation. In addition, p53 is increasingly being recognized as crucial player in nutrient sensing pathways. Hence, despite existence of contradictory findings and a varying density of evidence, several functions of p53 in adipocytes and ATs have been emerging, positioning p53 as an essential regulatory hub in ATs. Future studies need to make use of more sophisticated in vivo model systems and should identify an AT-specific set of p53 target genes and downstream pathways upon different (nutrient) challenges to identify novel therapeutic targets to curb metabolic diseases.
... Ortega et al. found an increase in p53 expression in the VAT of obese subjects and obese mice. The authors also showed that p53 expression is negatively correlated with IR [42]. Moreover, Huang et al. found a reduction in phosphorylated p53 in obese rats; according to these authors, p53 also inhibits Akt phosphorylation, which is needed for insulin signalling and GLUT4 translocation [1], [43]. ...
Article
Introduction: The aim of this is study was to analyse the expression of miR-193b, miR-378, miR-Let7-d, and miR-222 in human visceral adipose tissue (VAT), as well as their association with obesity, insulin resistance (IR), and their role in the regulation of genes controlling adipose tissue homeostasis, including adipocytokines, the phosphatase and tension homologue (PTEN), and tumour protein 53 (p53). Material and methods: VAT was obtained from normal-weight (NW), overweight, and obese (OW/OB) subjects with and without IR. Stem-loop RT-qPCR was used to evaluate miRNA expression levels. miRTarBase 4.0, miRWalk, and DIANA-TarBase v8 were used for prediction of validated target gene of the miRNA analysed. A qPCR was used to evaluate PTEN, p53, leptin (LEP), and adiponectin (ADIPOQ) mRNA. Results: miR-222 was lower in IR subjects, and miR-222 and miR-378 negatively correlated with HOMA-IR. PTEN and p53 are miR-222 direct targets according to databases. mRNA expression of PTEN and p53 was lower in OW/OB subjects with and without IR, compared to NW group and its levels positively associated with miR-222. Additionally, p53 and PTEN are positively associated with serum leptin levels. On the other hand, miR-193b and miR-378 negatively correlated with serum leptin but not with mRNA levels. Moreover, miR-Let-7d negatively correlated with serum adiponectin but not with adiponectin mRNA levels. Conclusions: Lower miR-222 levels are associated with IR, and PTEN and p53 expression; the implication of these genes in adipose tissue homeostasis needs more research.
... TP53 is the "guardian of the genome" and a major tumor suppressor gene, and in many types of cancer, major somatic mutations occur in TP53. The significance of p53 in heart disease, obesity, T2DM etc. has been demonstrated (45). Depending on the type of adipocytes, the roles of p53 on adipogenesis differ. ...
Article
Full-text available
Background Obesity is a medical problem with an increased risk for other metabolic disorders like diabetes, heart problem, arthritis, etc. Leptin is an adipose tissue-derived hormone responsible for food intake, energy expenditure, etc., and leptin resistance is one of the significant causes of obesity. Excess leptin secretion by poor diet habits and impaired hypothalamic leptin signaling leads to LR. Melatonin a sleep hormone; also possess antioxidant and anti-inflammatory properties. The melatonin can attenuate the complications of obesity by regulating its targets towards LR induced obesity.AimThe aim of this study includes molecular pathway and network analysis by using a systems pharmacology approach to identify a potential therapeutic mechanism of melatonin on leptin resistance-induced obesity.Methods The bioinformatic methods are used to find therapeutic targets of melatonin in the treatment of leptin resistance-induced obesity. It includes target gene identification using public databases, Gene ontology, and KEGG pathway enrichment by ‘ClusterProfiler’ using the R language, network analysis by Cytoscape, and molecular Docking by Autodock.ResultsWe obtained the common top 33 potential therapeutic targets of melatonin and LR-induced obesity from the total melatonin targets 254 and common LR obesity targets 212 using the data screening method. They are involved in biological processes related to sleep and obesity, including the cellular response to external stimulus, chemical stress, and autophagy. From a total of 180 enriched pathways, we took the top ten pathways for further analysis, including lipid and atherosclerosis, endocrine, and AGE-RAGE signaling pathway in diabetic complications. The top 10 pathways interacted with the common 33 genes and created two functional modules. Using Cytoscape network analysis, the top ten hub genes (TP53, AKT1, MAPK3, PTGS2, TNF, IL6, MAPK1, ERBB2, IL1B, MTOR) were identified by the MCC algorithm of the CytoHubba plugin. From a wide range of pathway classes, melatonin can reduce LR-induced obesity risks by regulating the major six classes. It includes signal transduction, endocrine system, endocrine and metabolic disease, environmental adaptation, drug resistance antineoplastic, and cardiovascular disease.Conclusion The pharmacological mechanism of action in this study shows the ten therapeutic targets of melatonin in LR-induced obesity.
... Strong evidence demonstrated the fundamental importance of p53 in metabolic diseases for example cardiovascular disease, obesity, and type 2 diabetes [32]. p53 is known as a negative regulator of adipogenesis in vitro, and also it was reported that p53 levels in white adipose tissue are augmented in diet-induced and genetic obesity mouse models and in obese humans [33]. ...
Article
Full-text available
Background Aging and obesity are the two major global health concerns. Sarcopenia, an age-linked disease, wherein a progressive loss of muscle volume, muscle strength, and physical activity occurs. In this study we evaluated the association of TP53 rs1625895 polymorphism with the susceptibility to sarcopenic obesity in Iranian old-age subjects. Methods Total of 176 old individuals (45 sarcopenic and 131 healthy) were recruited in this research and genotyped by PCR–RFLP. BMI, Skeletal Muscle Mass Index, body composition, Handgrip Strength, Gait Speed (GS), and biochemical parameters were measured. Chi-square test was done for genotypes and alleles frequency. Linear regression was applied to find the correlation between TP53 rs1625895 polymorphism, and biochemical and anthropometric parameters. The correlation between TP53 rs1625895 and the risk of sarcopenia and sarcopenic obesity was investigated by logistic regression. Results G allele was significantly higher in sarcopenic obesity group [ P = 0.037, OR (CI 95%) = 1.9 (1.03–3.5)] compared to A allele. BMI ( P = 0.049) and LDL (P = 0.04) were significantly differed between genotypes when GG was compared to AA/AG genotype. The results revealed when GG genotype compared to AA/AG genotype in adjusted model for age, the risk of sarcopenic obesity [ P value = 0.011, OR (CI 95%); 2.72 (1.25–5.91)] increased. Similarly, GG/AG genotype increased the risk of sarcopenic obesity [ P value = 0.028, OR (CI 95%); 2.43 (1.10–5.36)] in adjusted model for age compared to AA genotype. Conclusions We suggested that TP53 rs1625895 polymorphism may increase the risk of sarcopenic obesity in Iranian population.
... 27 Animal studies have also confirmed that R72 mice developed significantly increased fat accumulation and impaired insulin sensitivity. 18 Besides, increased p53 gene expression had been revealed in adipose tissue and endothelial cells of HFD-induced obese mice, 28 whereas HFD mice with endothelial cell-specific p53 deficiency showed improvement of insulin sensitivity and less fat accumulation. 29 In our study, we observed higher p53 levels in the hypothalami of HFD mice compared to those of controls, supporting the impact of p53 on the early puberty onset of HFD mice. ...
Article
Impact statement: High-fat intake and subsequent obesity are associated with premature onset of puberty, but the exact neuroendocrine mechanisms are still unclear. The transcriptional factor p53 has been predicted to be a central hub of the gene networks controlling the pubertal onset. Besides, p53 also plays crucial roles in metabolism. Here, we explored p53 in the hypothalami of mice fed a high-fat diet (HFD), which showed an up-regulated expression. Besides, we also revealed that overexpressed p53 may accelerate hypothalamo-pituitary-gonadal (HPG) axis activation partially through the c-Myc/Lin28/let-7 system. These results can deepen our understanding of the interaction between metabolic regulation and puberty onset control, and may shed light on the neuroendocrine mechanisms of obesity-related central precocious puberty.
Article
Reprogrammed cell metabolism is a well-accepted hallmark of cancer. Metabolism changes provide energy and precursors for macromolecule biosynthesis to satisfy the survival needs of cancer cells. The specific changes in different aspects of lipid metabolism in cancer cells have been focused in recent years. These changes can affect cell growth, proliferation, differentiation and motility through affecting membranes synthesis, energy homeostasis and cell signaling. The tumor suppressor p53 plays vital roles in the control of cell proliferation, senescence, DNA repair, and cell death in cancer through various transcriptional and non-transcriptional activities. Accumulating evidences indicate that p53 also regulates cellular metabolism, which appears to contribute to its tumor suppressive functions. Particularly the role of p53 in regulating lipid metabolism has gained more and more attention in recent decades. In this review, we summarize recent advances in the function of p53 on lipid metabolism in cancer. Further understanding and research on the role of p53 in lipid metabolism regulation will provide a potential therapeutic window for cancer treatment.
Article
Full-text available
p53 regulates several signaling pathways to maintain the metabolic homeostasis of cells and modulates the cellular response to stress. Deficiency or excess of nutrients causes cellular metabolic stress, and we hypothesized that p53 could be linked to glucose maintenance. We show here that upon starvation hepatic p53 is stabilized by O-GlcNAcylation and plays an essential role in the physiological regulation of glucose homeostasis. More specifically, p53 binds to PCK1 promoter and regulates its transcriptional activation, thereby controlling hepatic glucose production. Mice lacking p53 in the liver show a reduced gluconeogenic response during calorie restriction. Glucagon, adrenaline and glucocorticoids augment protein levels of p53, and administration of these hormones to p53 deficient human hepatocytes and to liver-specific p53 deficient mice fails to increase glucose levels. Moreover, insulin decreases p53 levels, and over-expression of p53 impairs insulin sensitivity. Finally, protein levels of p53, as well as genes responsible of O-GlcNAcylation are elevated in the liver of type 2 diabetic patients and positively correlate with glucose and HOMA-IR. Overall these results indicate that the O-GlcNAcylation of p53 plays an unsuspected key role regulating in vivo glucose homeostasis. p53 regulates signalling pathways involved in metabolic homeostasis. Here the authors show that O-GlcNAcylation of p53 in the liver plays a key role in the physiological regulation of glucose homeostasis, potentially via controlling the expression of the gluconeogenic enzyme phosphoenolpyruvate carboxykinase.
Article
Full-text available
Objective The purpose of this study was to investigate the relationships between TP53 Pro72Arg (rs1042522) polymorphism and susceptibility to type 2 diabetes (T2DM) and its related complications. Methods The TP53 Pro72Arg polymorphism was genotyped by polymerase chain reaction-restriction fragment length polymorphism (PCR-RFLP) method in 206 T2DM patients and 446 healthy controls. Mitochondrial DNA (mtDNA) content, mtDNA transcriptional level and large-scale mtDNA deletion were evaluated in leukocytes of T2DM patients using fluorescence-based quantitative PCR (FQ-PCR), reverse transcriptase-quantitative PCR (RT-qPCR) and long-range PCR approaches, respectively. The data of our study were processed by GraphPad Prism software (version 7.00). Results The distribution of TP53 Pro72Arg differed in T2DM patients from the controls, with a moderately increased proportion of TP53 Arg72 variant carriers (Pro/Arg and Arg/Arg genotypes) (88.3% vs 81.2%, p=0.022; OR=1.089, 95% CI=1.018–1.164). T2DM patients with Arg/Arg genotype had significantly decreased prevalences of diabetic neuropathy and retinopathy compared to those without (6.5% vs 19.4%, p=0.018 and 14.8% vs 30.7%, p=0.018, respectively). T2DM patients with Arg/Arg genotype had higher mtDNA content and mtRNA expression level than those who were not Arg/Arg genotype (p<0.05 for all), and we did not observe mtDNA 4977-base pair (bp) deletion mutations in the leukocytes of T2DM patients. Conclusion There was a significant association of the TP53 Pro72Arg polymorphism with susceptibility to T2DM, and the homozygous Arg/Arg genotype of this gene locus might be a protective factor for diabetic complications. Those results suggested that the TP53 Arg72 variant had a different association with type 2 diabetes and its complications, and it might be related to mtDNA maintenance of the TP53 Arg72 variant under hyperglycemia-induced stress.
Article
Obesity has emerged as a leading cause of death in the last few decades, mainly due to associated cardiovascular diseases. Obesity, inflammation, and insulin resistance are strongly interlinked. Lisofylline (LSF), an anti-inflammatory agent, demonstrated protection against type 1 diabetes, as well as reduced obesity-induced insulin resistance and adipose tissue inflammation. However, its role in mitigating cardiac inflammation associated with obesity is not well studied. Mice were divided into 4 groups; the first group was fed regular chow diet, the second was fed regular chow diet and treated with LSF, the third was fed high fat diet (HFD), and the fourth was fed HFD and treated with LSF. Cardiac inflammation was interrogated via expression levels of TNF α, interleukins 6 and 10, phosphorylated STAT4 and lipoxygenases 12 and 12/15. Apoptosis and expression of the survival gene, AMPK, were also evaluated. We observed that LSF alleviated obesity-induced cardiac injury indirectly by improving both pancreatic β-cell function and insulin sensitivity, as well as, directly via upregulation of cardiac AMPK expression and downregulation of cardiac inflammation and apoptosis. LSF may represent an effective therapy targeting obesity-induced metabolic and cardiovascular complications.
Article
Full-text available
Roux-en-Y gastric bypass and biliopancreatic diversion can markedly ameliorate diabetes in morbidly obese patients, often resulting in disease remission. Prospective, randomized trials comparing these procedures with medical therapy for the treatment of diabetes are needed. In this single-center, nonblinded, randomized, controlled trial, 60 patients between the ages of 30 and 60 years with a body-mass index (BMI, the weight in kilograms divided by the square of the height in meters) of 35 or more, a history of at least 5 years of diabetes, and a glycated hemoglobin level of 7.0% or more were randomly assigned to receive conventional medical therapy or undergo either gastric bypass or biliopancreatic diversion. The primary end point was the rate of diabetes remission at 2 years (defined as a fasting glucose level of <100 mg per deciliter [5.6 mmol per liter] and a glycated hemoglobin level of <6.5% in the absence of pharmacologic therapy). At 2 years, diabetes remission had occurred in no patients in the medical-therapy group versus 75% in the gastric-bypass group and 95% in the biliopancreatic-diversion group (P<0.001 for both comparisons). Age, sex, baseline BMI, duration of diabetes, and weight changes were not significant predictors of diabetes remission at 2 years or of improvement in glycemia at 1 and 3 months. At 2 years, the average baseline glycated hemoglobin level (8.65±1.45%) had decreased in all groups, but patients in the two surgical groups had the greatest degree of improvement (average glycated hemoglobin levels, 7.69±0.57% in the medical-therapy group, 6.35±1.42% in the gastric-bypass group, and 4.95±0.49% in the biliopancreatic-diversion group). In severely obese patients with type 2 diabetes, bariatric surgery resulted in better glucose control than did medical therapy. Preoperative BMI and weight loss did not predict the improvement in hyperglycemia after these procedures. (Funded by Catholic University of Rome; ClinicalTrials.gov number, NCT00888836.).
Article
Full-text available
We are increasingly aware that cellular metabolism plays a vital role in diseases such as cancer, and that p53 is an important regulator of metabolic pathways. By transcriptional activation and other means, p53 is able to contribute to the regulation of glycolysis, oxidative phosphorylation, glutaminolysis, insulin sensitivity, nucleotide biosynthesis, mitochondrial integrity, fatty acid oxidation, antioxidant response, autophagy and mTOR signalling. The ability to positively and negatively regulate many of these pathways, combined with feedback signalling from these pathways to p53, demonstrates the reciprocal and flexible nature of the regulation, facilitating a diverse range of responses to metabolic stress. Intriguingly, metabolic stress triggers primarily an adaptive (rather than pro-apoptotic) p53 response, and p53 is emerging as an important regulator of metabolic homeostasis. A better understanding of how p53 coordinates metabolic adaptation will facilitate the identification of novel therapeutic targets and will also illuminate the wider role of p53 in human biology.
Article
Tumorigenesis is associated with enhanced cellular glucose uptake and increased metabolism. Because the p53 tumor suppressor is mutated in a large number of cancers, we evaluated whether p53 regulates expression of the GLUT1 and GLUT4 glucose transporter genes. Transient cotransfection of osteosarcoma-derived SaOS-2 cells, rhabdomyosarcoma-derived RD cells, and C2C12 myotubes with GLUT1-P-Luc or GLUT4-P-Luc promoter-reporter constructs and wild-type p53 expression vectors dose dependently decreased both GLUT1 and GLUT4 promoter activity to approximately 50% of their basal levels. PG13-Luc activity, which was used as a positive control for functional p53 expression, was increased up to ∼250-fold by coexpression of wild-type p53. The inhibitory effect of wild-type p53 was greatly reduced or abolished when cells were transfected with p53 with mutations in amino acids 143, 248, or 273. A region spanning −66/+163 bp of the GLUT4 promoter was both necessary and sufficient to mediate the inhibitory effects of p53. Furthermore, in vitro translated p53 protein was found to bind directly to two sequences in that region. p53-DNA binding was completely abolished by excess unlabeled probe but not by nonspecific DNA and was super-shifted by the addition of an anti-p53 antibody. Taken together, our data strongly suggest that wild-type p53 represses GLUT1 and GLUT4 gene transcription in a tissue-specific manner. Mutations within the DNA-binding domain of p53, which are usually associated with malignancy, were found to impair the repressive effect of p53 on transcriptional activity of the GLUT1 and GLUT4 gene promoters, thereby resulting in increased glucose metabolism and cell energy supply. This, in turn, would be predicted to facilitate tumor growth.
Article
Adipose tissue inflammation induces systemic insulin resistance in persons with obesity and heart failure, and has a crucial role in the progression of these diseases. Chronic inflammatory processes share a common mechanism in which increased production of reactive oxygen species activates p53 and NF-κB signaling, leading to up-regulation of pro-inflammatory cytokine expression and impairment of glucose metabolism. Since inhibition of these processes could slow the progression of various diseases, targeting adipose inflammation has the potential to become a new therapeutic approach for diabetes and heart failure.
Article
Objectives: Tumor protein p53 is a transcription factor involved with cellular responses to stressors including limited glucose availability. We hypothesized that modulating p53 levels would affect cellular glucose uptake. Methods and results: Transfecting cultured primary mouse hepatocytes with p53 siRNA suppressed p53 mRNA expression >90%. Control hepatocytes (transfected with non-targeting siRNA) increased glucose uptake (2.28 ± 1.02-fold vs basal, p 0.009) in response to 100 nM insulin, but p53 siRNA-treated hepatocytes had a blunted response (0.92 ± 0.11-fold vs basal; between group difference p 0.0012). In adipocytes differentiated from the pre-adipocyte line 3T3-L1, knockdown of p53 had no effect on insulin-stimulated glucose uptake. There were no differences in Glut 1 or Glut 2 expression in the plasma membrane fraction or in the levels of phosphorylated AKT in cell lysates between primary hepatocytes transfected with p53 siRNA or control siRNA. Glycemic responses to insulin tolerance, glucose tolerance, and pyruvate tolerance tests did not differ between p53 knockout and wild type mice. Discussion: Thus, inhibition of p53 has pleiotropic effects, inhibiting glucose uptake in the liver but having no effect on adipocytes. Knockout of p53 has no apparent effect on glucose homeostasis in intact lean mice. An explanation for the association between p53 expression and hepatocyte glucose uptake remains to be elucidated.
Article
The in vitro and in vivo anti-melanoma effect of antidiabetic drug metformin was investigated using B16 mouse melanoma cell line. Metformin caused a G(2)/M cell cycle arrest associated with apoptotic death of melanoma cells, as confirmed by the flow cytometric analysis of cell cycle/DNA fragmentation, phosphatidylserine exposure and caspase activation. Metformin-mediated apoptosis of melanoma cells was preceded by induction of oxidative stress and mitochondrial membrane depolarization, measured by flow cytometry in cells stained with appropriate fluorescent reporter dyes. The expression of tumor suppressor protein p53 was increased, while the mRNA levels of anti-apoptotic Bcl-2 were reduced by metformin, as revealed by cell-based ELISA and real-time RT-PCR, respectively. Treatment with metformin did not stimulate expression of the cycle blocker p21, indicating that p21 was dispensable for the observed cell cycle arrest. The activation of AMP-activated protein kinase (AMPK) was not required for the anti-melanoma action of metformin, as AMPK inhibitor compound C completely failed to restore viability of metformin-treated B16 cells. Metformin induced autophagy in B16 cells, as demonstrated by flow cytometry-detected increase in intracellular acidification and immunoblot-confirmed upregulation of autophagosome-associated LC3-II. Autophagy inhibitors ammonium chloride and wortmannin partly restored the viability of metformin-treated melanoma cells. Finally, oral administration of metformin led to a significant reduction in tumor size in a B16 mouse melanoma model. These data suggest that anti-melanoma effects of metformin are mediated through p21- and AMPK-independent cell cycle arrest, apoptosis and autophagy associated with p53/Bcl-2 modulation, mitochondrial damage and oxidative stress.
Article
Background  Obesity is associated with premature mortality, particularly when very severe and/or complicated by significant co-morbidities such as diabetes mellitus, cardiovascular and respiratory disease. Conventional management of obesity, namely diet, exercise, behavioural modification and pharmacotherapy has limited and poorly sustained effects on weight loss and uncertain benefits for survival. Objectives  We aimed to review the literature in to determine whether bariatric surgery for morbid obesity prolongs life. Search Strategy  A Search was conducted of data bases including Medline, Cochrane library, and Science Direct. Results  Bariatric surgery produces significant and sustained weight loss. Greater weight loss occurs with procedures that have both a restrictive and malabsorptive component. In addition to resolution of, or at least improvement in co-morbidities and enhanced quality of life, six studies provide compelling evidence that bariatric surgery significantly prolongs life, an effect which is most marked in diabetics and predominantly attributable to reductions in death due to cardiovascular causes and cancer. Conclusion  Taken together, the cost, quality of life, and survival benefits provide a compelling argument for the provision of bariatric surgery as a management strategy of choice for severe obesity, particularly when associated with diabetes mellitus, or other factors conferring a significant cardiovascular risk. The optimal procedure and strategy for patient selection remains to be determined.