ArticlePDF AvailableLiterature Review

The role of adiponectin signaling in metabolic syndrome and cancer

Authors:

Abstract and Figures

The increased prevalence of obesity has mandated extensive research focused on mechanisms responsible for associated clinical complications. Emerging from the focus on adipose tissue biology as a vitally important adipokine is adiponectin which is now believed to mediate anti-diabetic, anti-atherosclerotic, anti-inflammatory, cardioprotective and cancer modifying actions. Adiponectin mediates these primarily beneficial effects via direct signaling effects and via enhancing insulin sensitivity via crosstalk with insulin signaling pathways. Reduced adiponectin action is detrimental and occurs in obesity via decreased circulating levels of adiponectin action or development of adiponectin resistance. This review will focus on cellular mechanisms of adiponectin action, their crosstalk with insulin signaling and the resultant role of adiponectin in cardiovascular disease, diabetes and cancer and reviews data from in vitro cell based studies through animal models to clinical observations.
No caption available
… 
Content may be subject to copyright.
The role of adiponectin signaling in metabolic syndrome
and cancer
Michael P. Scheid & Gary Sweeney
#
Springer Science+Business Media New York 2013
Abstract The increased prevalence of obesity has mandated
extensive research focused on mechanisms responsible for
associated clinical complications. Emerging from the focus
on adipose tissue biology as a vitally important adipokine is
adiponectin which is now believed to mediate anti-diabetic,
anti-atherosclerotic, anti-inflammatory, cardioprotective and
cancer modifying actions. Adiponectin mediates these primar-
ily beneficial effects via direct signaling effects and via en-
hancing insulin sensitivity via crosstalk with insulin signaling
pathways. Reduced adiponectin action is detrimental and oc-
curs in obesity via decreased circulating levels of adiponectin
action or development of adiponectin resistance. This review
will focus on cellular mechanisms of adiponectin action, their
crosstalk with insulin signaling a nd the resultant role of
adiponectin in cardiovascular disease, diabetes and cancer
and reviews data from in vitro cell based studies through
animal models to clinical observations.
Keywords Adiponectin
.
Signaling
.
Metabolism
.
Diabetes
.
Cardiovascular disease
.
Cancer
1 Introduction
1.1 Introduction to adiponectin physiology
Due to numerous factors, such as an increase in sedentary
lifestyles combined with poor diet and/or over-nutrition, the
prevalence of overweight and obesity has escalated dramati-
cally and brought with it increases in various associated dis-
ease states [1, 2 ]. For example, it is clear that obesity is
associated with an increased risk for development of type 2
diabetes and var ious forms of cardiovascular disease [2].
Recently, significant emphasis has been placed on associa-
tions between obesity and various forms of cancer and strong
relationships occur [1]. Therefore, the search for mechanisms
responsible for cancer, cardiovascular disease and diabetes in
obesity has become intensive and it is becoming clear that
many answers may lie in the adipose tissue itself. We now
know that as well as fulfilling its traditional role as a site for
storage of excess energy, adipose tissue is an important endo-
crine organ. A large array of adipose-derived factors (collec-
tively termed adipokines) has now been discovered and their
effects characterized [3]. Whereas the majority of adipokines,
such as tumor necrosis factor-α, adipocyte fatty acid binding
protein (A-FABP) and lipocalin-2 are proinflammatory,
adiponectin is one adipokine that has been shown to have
anti-inflammatory, anti-diabetic, cardioprotective and cancer-
altering actions [46]. Dysfunctional adipose tissue in obesity
releases a disturbed profile of adipokines with elevated levels
of proinflammatory factors and reduced adiponectin. This
review will focus on adiponectin signaling mechanisms and
the implications of altered adiponectin action in metabolic
syndrome and cancer.
1.2 Crosstalk between adiponectin and insulin signaling
Adiponectin binds to two distinct, seven-transmembrane re-
ceptors (AdipoR1 and AdipoR2) that share 67 % amino acid
identity [7]. AdipoR2 differs significantly with AdipoR1 near
the amino-terminus that resides in the cytoplasm of the cell,
and appears to conduct the majority of the signaling by each
receptor through association with downstream effector mole-
cules. AdipoR1 is expressed ubiquitously and highest in skel-
etal muscle, while AdipoR2 is expressed predominantly in the
liver [7]. The adiponectin receptors communicate to intracel-
lular pathways through the activation of serine/threonine ki-
nases such as AMPK and AKT2, increased phospholipase C
activity, and small G-proteins including Rab5, which will be
M. P. Scheid
:
G. Sweeney (*)
Department of Biology, York University, Toronto,
ON M3J 1P3, Canada
e-mail: gsweeney@yorku.ca
Rev Endocr Metab Disord
DOI 10.1007/s11154-013-9265-5
described in further detail below. Cloning of the adiponectin
receptor [7] and its genetic ablation have accelerated under-
standing of how the adiponectin receptor couples to these
pathways (Fig. 1).
An important function of the adiponectin receptor is to
sensitize and increase the magnitude of insulin signaling on
target cells, facilitating glucose uptake and energy homeostasis
[8]. One way in which adiponectin sensitizes target tissues to
insulin is through the increase in 5-adenosine monophosphate
kinase (AMPK) activity, whic h stimulates fatty-acid oxidation,
increases PPARa expression, and increases glucose uptake,
while repressing gly colysis, lipid biogenesis and gluconeogen-
esis (reviewed in [9]). In addition, AMPK can directly increase
insulin sensitivity by stimula ting the phosphorylation of perox-
isome proliferator-activated receptor-γ co-activator 1alpha
(PGC-α [10]), a transcription co-activator that plays a crit-
ical role in the biosynthesis of mitochondria and oxidative
phosphorylation. The deacetylase SIR T1 also acts on PGC-α
[11, 12] and both AMPK and SIRT1 appear to act coopera-
tively to fully activate PGC-α [13].
AMPK is a heterotrimeric protein composed of α, β and γ
subunits that stabilize and activate the protein kinase upon
interaction of AMP to Bateman repeats of the γ subunit [14].
An upstream kinase, termed AMPKK, phosphorylates the α
subunit at threonine-172 within the activation loop, which
increases catalytic activity. The primary AMPKK has been
identified as a complex composed of liver kinase B1 (LKB1),
together with two allosteric regulators MO25 and STRAD
[1517]. CaMKKβ and the transforming growth factor beta-
activated kinase-1 (TAK1) can also serve as AMPKKs under
some conditions [1821], allowing multiple inputs for activa-
tion depending on the cell type and stimulus. In addition to its
well established role in energy homeostasis, AMPK has also
recently been identified as a tumor suppressor by virtue of its
opposition to aerobic glycolysis [22, 23], and loss of AMPK
or the LKB1-STRAD-MO25 complex accelerates tumorigen-
esis for many human cancers, as will be described below.
Adiponectin can activate AMPK th rough an LKB1-
STRAD-MO25 dependent mechanism, as well as CaMKKβ
mechanism in muscle [2431]. The mechanism of activation
requires the adapter protein APPL1 [28, 32], which binds to
the N-terminus of the AdipoR1 [28]. APPL1 is composed of a
C-terminal PTB domain, an N-terminal Bin/amphiphysin/Rvs
(BAR) domain, and a central pleckstrin homology (PH) do-
main [32
]. Each domain serves as docking sites for down-
stream effectors, and deletion of each has discrete outcomes;
for example, deletion of the PTB domain disrupts binding to
AdipoR1, while deletion of the BAR domain of APPL1
disrupts binding with LKB1 [28].
Mechanistically, LKB1 is a nucleo-cytoplasmic shuttling
protein [33], and binding to APPL1 could act to shift the
localization of LKB1 to the cytoplasm, thereby increasing
the colocalization of LKB1 with AMPK [31]. An important
question that remains is how adiponectin facilitates APPL1
LKB1 interaction, and why it is not maintained constitutively
in untreated cells. It is possible that other signals modify either
Fig. 1 Adiponectin sensitizes
Insulin signaling by activating the
metabolic checkpoint kinase
AMPK through APPL1-mediated
recruitment of LKB and
CaMKK2, which increases fatty
acid oxidation through activation
of PPARa. AKT2 associates with
APPL1 and is synergistically
activated by adiponectin and
insulin. AMPK also activates the
GTPase TSC1/TSC2 complex,
which acts on the small G-protein
Rheb. This promotes inactivation
of the mTOR complex 1
(TORC1), suppressing protein
synthesis and shifts the cell away
from aerobic glycolysis. Red
arrows indicate an inhibitor effect
Rev Endocr Metab Disord
APPL1 or LKB1 during adiponectin treatment, which then
directs APPL1LKB1 association. Zhou and colleagues re-
ported that phospholipase C activity was stimulated by
adiponectin, resulting in intracellular Ca
+2
release [31]. It is
possible that a Ca
+2
activated kinase, such as PKC, could
phosphorylate and alter APPL1. One study did investigate
the potential post-translational modifications of APPL1,
where it was shown to undergo PKCα-mediated phosphory-
lation of Ser430. Use of the phosphorylation prediction pro-
gram Scansite (http://scansite.mit.edu) reveals that APPL1 has
several other potential sites, including Thr177 that lies directly
within the BAR domain and could be targeted by an AGC
kinase such as PKC. Future studies might test whether Thr177
is a phosphorylation site and if this alters LKB1 association. In
addition to APPL1, LKB1 has recently been shown to also
undergo modification following adiponectin stimulation.
Ser307 is required for nuclear localization of LKB1 through
aPKCζ dependent mechanism [34], and phosphorylation of
this site is reduced following adiponectin stimulation [35],
resulting in its cytoplasmic translocation. The mechanism of
dephosphorylation of LKB1 is suggested to be as a result of
reduction in PKCζ activity due to PP2A-mediated dephos-
phorylation [35]. Thus, adiponectin reduces PKCζ activity,
resulting in a shift in the rate of LKB1 nuclear localization.
APPL1 also binds to the protein kinase AKT2 and the cata-
lytic subunit of phosphoinositide-3-kinase (PI3K) [32].
Activation of AKT2 requires co-localization to the plasma mem-
brane, where binding of its PH domain to phosphatidylinositol-
3,4,5-P
3
(PIP
3
) results in a conformation change, relieving auto-
inhibition and allowing accessibility by its upstream kinases
PDK1 and mTOR-containing complex-2 (TORC2; reviewed in
[36]). These kinases phosphorylate AKT2 at Thr309 and Ser474,
respectively, resulting in full activation [37]. PIP
3
is generated by
PI3K, which is activated through binding with phosphotyrosine
sites on the insulin receptor substrate-1 (IRS-1). Since APPL1
binds to AKT2, this could facilitate co-localization with insulin
receptors to amplify the AKT2 driven signal. There is evidence
that this mechanism occurs, as several studies have shown that
adiponectin can synergistically activate AKT2 through an
APPL1-dependent mechanism [28, 38, 39], thereby promoting
glucose uptake and the activation of pro-surv ival pathway in
some cell types, leading to enhanced survival of cells following
stress insu lts. APPL1 als o facilitates Glut4 and insulin receptor
translocation to the plasma membrane through a mechanism that
involves APPL1-dependent activation of the Rab5 G-protein
[28, 40]. In yet another mechanism, APPL1 has been shown to
regulate the interaction between AKT and the AKT inhibitor
protein tribble 3 (TRB3). Overexpression of APPL1 disrupts
AKT-TRB3 interaction, and knockdown of APPL1 augmented
this interaction [41].
In addition to these many associations, APPL1 also binds
with the catalytic domain of PI3K, which may also have
implications on AKT2 activation. Although adiponectin is not
classically thought to stimulate the PI3K pathway, APPL1
could act as a scaffold to facilitate PI3K activation following
tyrosine phosphorylation of IRS1 by the insulin receptor. Since
engagement of the SH2 domains of the p85 subunit of PI3K is
necessary for the activation of PI3KCA, this mechanism would
ensure that adiponecti n could only contribute to AKT2 signal-
ing during periods of insulin stimulation. Consistent with this
model, inhibition o f PI3K with the inhibitors wortmannin or
LY294002 appear to block signaling by the adiponectin recep-
tor [42], suggesting tha t co-localizatio n of PI3K could play an
impo rtant role. Furthermore, expression of AdipoR1 and
AdipoR2 are influenced by PI3K-dependent transcription, in
which AKT phosphorylation of the transcription factor FOXO1
represses ADIPOQR gene activation [43], suggesting a nega-
tive feedback mechanism to limit sensitization.
Finally, adiponectin may also sensitize target cells to insulin
by repressing feedback inhibition signals at the level of the
insulin receptor. For example, growth and anti-apoptotic signals
are induced by tyrosine phosphorylation of IRS1, which binds to
and activates PI3K [44], which in turn activates AKT. This
pathway leads to inactivation of the GTPase complex of TSC1
and TSC2, allowing GTP loading of the small G-protein ras-
homolog enriched in brain (Rheb). Rheb is needed for mTOR
activation within mTOR complex 1 (TORC1), a multiprotein
complex that coordinates mTOR activity, substrate specificity ,
and sensitivity to the macrolide rapamycin [45]. TORC1 phos-
phorylates the p70 S6 ribosomal protein kinase (S6K) within the
hydrophobic motif at Thr389, resulting in full activation, and so
a linear pathway exists that leads from the insulin receptor to
activation of S6K and TORC1. Once activated, TORC1 and
S6K target components of the protein translation machinery and
mitochondrial biogenesis, resulting in cell growth [46].
Following insulin stimulation, IRS1 signaling is negatively
regulated by S6K in a classic feedback inhibition loop [47,
48]. Several groups have shown that adiponectin can promote
increased insulin/IGF-1 signaling by down regulating S6K
activity, which in turn results in the loss of a negative phos-
phorylation event on IRS1, thereby reducing negative feed-
back inhibition [25, 49]. Adiponectin may also increase IRS1
phosphorylation within the YXXM motif at Y612, leading to
binding of the SH2 domain of p85 and activation of PI3K and
AKT [49]. Together these observations are consistent with the
finding that adiponectin can activate the pro-survival PI3K/
AKT pathway in some cell types [38, 39
], leading to survival
of muscle and hepatocytes following stress insults.
1.3 Adiponectin signaling and regulation of metabolism
in cardiovascular disease
Obesity and the associated metabolic syndrome make individ-
uals more inclined to develop cardiovascular dysfunctions
which can be devastating owing to the high risk for mortality
or loss of quality of life [50, 51]. Accordingly, there is strong
Rev Endocr Metab Disord
interest in determining the various ways via which obesity can
influence initiation and progression of cardiovascular disease.
Adiponectin has widespre ad effe cts on cardiovascu lar disease,
and here we will focus on insulin signaling mechanisms regu-
lating direct effects of adiponectin on the heart and also the
vasculature [5, 52].
A large number of studies in adiponectin knockout (Ad-
KO) mice has established that lack of adiponectin exacerbates
stress-induced remodeling events such as hypertrophy, apo-
ptosis and fibrosis and that these can be prevented or corrected
by replenishing adiponectin to these animals [5358]. Most
importantly, changes in cardiac metabolism are central to the
pathogenesis of heart failure in obesity [2, 5961]and
adiponectin has been shown to act as an orchestrator of cardiac
fatty acid and glucose uptake and metabol ism. APPL1-
dependent activation of AMPK plays a critical role in medi-
ating cardiometabolic effects of adiponectin [62, 63]. In pri-
mary rat cardiomyocytes adiponectin increased association of
AdipoR1 with APPL1, subsequent binding of APPL1 with
AMPKα2 then phosphorylation and inhibition of ACC [62].
The ensuing increased fatty acid uptake and oxidation could
be attenuated by using siRNA to efficiently knockdown
APPL1 [62]. We have unpublished observations that APPL1
transgenic mice are protected from high fat diet induced
cardiomyopathy. Adiponectin also acts via APPL1 to protect
cardiomyocytes from hypoxia/reoxygenation-induced apo-
ptosis [64]. APPL2 can bind APPL1, sequestering it to pre-
vent the interaction with AdipoR in muscle cells and thus acts
as a negative regulator of adiponectin, and insulin, signaling
[65]. However, little is known yet regarding the role of APPL2
in the heart. In addition to activating AdipoR1-APPL1-
AMPK signaling, adiponectin enhanced insulin-stimulated
Akt phosphorylat ion leading to glucose uptake in primary adult
rat c ardiomyocy tes [62]. Thus, adiponectin enhan ces glucose
uptake via both insulin-mim etic and insulin-sensitizing
mechanisms.
Studies such as those described abov e demonstrating
cardioprotective effects of adipoenctin have been largely sup-
ported by numerous clinical studies which established inverse
correlations between plasma adiponectin levels and occurrence
or severity of heart failure [52]. However, the cardioprotective
role of adiponectin is not without question since various studies
have now found positive correlations between adiponectin
levels and adverse outcome in certain individuals. In particular,
elevated adiponectin levels have been found in late stages of
chronic heart failure and this has been suggested to reflect either
a permissive role of adiponectin or a compensatory increase
designed to overcome the development of adiponectin resis-
tance [66]. The latter is of particular interest and one potential
mechanism of adiponectin resistance is loss of one or more
receptor isoforms. Previous studies have indeed reported that
AdipoR expression was suppressed concomitantly with elevat-
ed adiponectin levels in failing hearts compared with control
subjects, and increased back toward normal levels after me-
chanical unloading [67]. In another study, neither AdipoR1 or
AdipoR2 were found to be altered in hearts from patients with
dilated cardiomyopathy [68]. Cardiac AdipoR1 expression also
decreased early after onse t of streptozotocin-induced diabetes
yet increased after longer duration of diabetes [69] and cardiac
AdipoR2 levels were decreased in diabetic rats [70].
Hyperglycemia or hyperinsulinemia which are common in
many obese individuals can both directly regulate AdipoR
isoform expression [71, 72
]. In addition to correlative studies,
functional tests have shown that a higher dose of adiponectin
was required to blunt oxidative and nitrative stress in hearts of
mice after high fat diet feeding [73]. The well estabilshed effect
of adiponectin to protect against hypoxia reoxygenation in-
duced oxidative stress and cell death was lost using cells
derived from AdipoR1-KO mice [74]. Whether adiponectin
resistance develops and how this occurs must now be more
clearly understood.
Adiponectin signaling in the vasculature also has signifi-
cant impact in vascular function and will be discussed briefly
here. First, adiponectin mediates several effects on the endo-
thelium to combat endothelial dysfunction and the majority of
these depend upon AMPK-mediated signaling. Both
AdipoR1 and AdipoR2 are expressed in endoth elial cells
and adiponectin activates AMPK-dependent phosphorylation
of eNOS to produce nitric oxide and induce vasodilation.
Knockdown of APPL1 expression reduced adiponectin-
stimulated phosphorylation of eNOS at Ser1177, and the
association of eNOS and heat shock protein 90 [75]. These
events resulted in reduced NO production. Adiponectin also
mediates anti-inflammatory effects on endothelium via PKA-
dependent suppression of NF-κB acti vation [76, 77].
Additional beneficial endothelial effects of adiponctin include
attenuating ROS production and decreasing monocyte attach-
ment which, together with inhibition of smooth muscle pro-
liferation and migration contribute to prevention of athero-
sclerosis [5]. A more recent interesting aspect of adiponectins
regulation of vascular function is the ability to mobilize and
increase function of endothelial progenitor cells. The phenom-
enon of adiponectin resistance may also pertain to vascular
effects since the expression of APPL1 was reduced in vessels
from various commonly used obese and diabetic animal
models [75, 78, 79].
1.4 Adiponectin signaling and regulation of metabolism
in diabetes
The substantial evidence linking adiponectin to metabolic
syndrome has been extensively reviewed [4, 80]. AdipoR1
and AdipoR2 appear to be the main adiponectin receptor
isoforms involved in mediating the metabolic effects of
adiponectin [80, 81] although there is also evidence that T-
cadherin may mediate or facilitate signaling [82, 83]. In
Rev Endocr Metab Disord
skeletal muscle, adiponectin stimulates glucose uptake via
inducing translocation of the glucose transporter GLUT4 to
the cell surface [84] and this effect was abolished upon knock-
down of APPL1 using siRNA [28]. An important downstream
target of AdipoR/APPL1 signaling in mediating metabolic
effects is AMPK [80, 81]. Adiponectin also enhances insulin
sensitivity and dominant negative AMPK overexpression in
muscle reduced the insulin sensitizing action of adiponectin
[25, 85]. AdipoR-APPL1-AMPK signaling has also been
shown to be important in mediating hepatic metabolic effects
of adiponectin [86].
Various mouse models have been informative in determin-
ing the metabolic consequences of ad iponec tin signaling
patways. First, adiponectin knockout mice have a relatively
normal phenotype until challenged by a stress such as high fat
diet, under which conditions they tend to show exacerbated
insulin reistance and metabolic dysfunction [8789]. Deletion
of both AdipoR1 and R2 in mice led to increased lipid accu-
mulation in various tissues, insulin resistance and glucose
intolerance [ 90]. AdipoR1 deletion results in lack of
adiponectin-stimulated AMPK activation and upon AdipoR2
deletion the principal signaing defect occurs in PPARα sig-
naling [90]. AdipoR1 KO mice exhbited decreased glucose
tolerance and defects in AMPK activation [90, 91]whereas
young AdipoR2 KO mice were found to have improved
glucose tolerance and lipid profiles in response to high fat diet
yet older mice had worse metabolic dysfunction [91]. Deletion
of AdipoR1 specifically in skeletal muscle led to defects in
peripheral glucose homeostasis [24]. Introduction of APPL1
to skeletal muscle in rats by electrotransfer-mediated
overexpression enhanced insulin-stimulated glucose disposal
[92]. Recently, mice lacking the APPL1 gene were generated
and phenotypically normal, and isolated embryonic fibro-
blasts stimulated with insulin showed no change in activation
of AKT1 or AKT2 [93]. Whether these results are a result of
compensatory effects involving APPL2 or some other protein
as a result of chronic depletion of APPL1 are not clear.
Adiponectin enhanced skeletal muscle biogenesis and a fiber
type switch to provide more oxidative capacity through an
AMPK-myocyte enhancer factor 2C (MEF2C)-PGC1α path-
way [24, 94, 95].
Adiponectin resistance is also likely to be a physiologically
relevant phenomenon in peripheral metabolic tissues and in-
deed insulin may be one important regulator of peripheral
adiponectin sensitivity. Providing insulin to streptozotocin-
induced diabetic mice reduced the elevated levels of AdipoR
expression found in liver and skeletal muscle of these animals
[43]. Insulin treatment of L6 skeletal muscle cells decreased
AdipoR1 levels while contrasting effects have been reported
on AdipoR2 [43, 72]. In L6 cells high glucose treatment also
reduced AdipoR1 and AdipoR2 expression in L6 cells [72].
Exercise can influence AdipoR expression in various tissues
[96] and more recent studies have shown increased expression
of APPL1 in response to exercise [
97, 98]. APPL1 ex-
pression and phosphorylation on Ser401 were elevated in obese
individuals with type 2 diabetes and reduced after bariatric
surge ry [99].
1.5 Adiponectin signaling and regulation of metabolism
in cancer
In addition to metabolic disease, the importance of nutrient
and energy homeostasis is becoming increasingly evident for
other pathologies such as cancer, where a fine balance is
achieved to optimize cell division, evade apoptosis, and es-
cape the localized stromal environment to metastasize to
distant organs [100, 101]. Furthermore, cancer is a multifac-
torial disease characterized by transformation-dependent ge-
nomic lesions, which act in cooperation with both the system-
ic environment and the tumor microenvironment. Importantly,
circulating and localized hormones and other soluble regula-
tors produced by adipose tissue can both positively and neg-
atively alter tumor growth. These effects could include alter-
ations of tumor physiology, immune system activation, pro-
moting angiogenesis, and by influencing metastasis to distant
organs [102]. Since caloric intake and obesity are associated
with cancer risk [6, 103105], it is likely that circulating
hormones produced by adipocytes, such as adiponectin, play
a significant role in malignancies [6]. In particular, breast
cancer is uniquely positioned t o interact with adipose-
derived factors since breast tissue is predominantly adipose
tissue, providing an adipocytokine microenvironment that
interacts with the evolving tumor mass.
In the context of cancer initiation and acceleration, obesity
and insulin resistance correlates positively with tumor growth
and poor prognosis. Under these same conditions, circulating
adiponectin levels are low, suggesting that adiponectin could
therefore have anti-cancer actions [106, 107]. Direct correlations
between adiponectin levels and breast cancer incidence have
been recorded by numerous studies, with all demonstrating a
moderate to strong association between lower serum adiponectin
levels and breast cancer risk [108112]. Furthermore, several
in vitro studies have examined cultured breast cancer cell lines
and their response to exogenously added adiponectin [27,
113121], and there appears to be a consistent repressive effect
on growth, adhesion and migration caused by adiponectin. A
limitation of these studies, however, is the use of transformed,
aneuploidic cell lines in all cases. In addition, the interac tion
between tumor cell and the stroma is absent and therefore other
signaling factors that could crosstalk and alter the genetic profile
of adiponectin-induced gene expression are lost. Finally , serum
levels of adiponectin may not represent the localized microenvi-
ronment of the breast tumor or the predominant adipocyte-
derived stroma, confounding observations made solely on circu-
lating concentrations.
Rev Endocr Metab Disord
One genetic study performed with mice showed that re-
duced expression of adiponectin could lead to accelerated
breast cancer initiation [12 2]. These authors utilized
MMTV-driven polyomavirus middle T antigen transgenic
mice within an adiponectin haplodeficient background.
Tumors grew faster, were more aggressive, and displayed
increased activation of PI3K/AKT. Inactivation of PTEN in
these tumors was noted and occurred as a result of thioredoxin
conjugation, suggesting that reduced adiponectin promotes
tumorigenesis by promoting thioredoxin adducts [122]. It
remains to be determined if this is a general mechanism that
also occurs in human breast tumors.
1.6 Adiponectin inhibition of the Warburg effect
It has been known for years that tumor cells become
addicted to aerobic glycolysis as a primary source of energy
and biosynthetic materials, in a process termed the Warburg
effect for its first observation by Otto Warburg more than
80 years ago (reviewed in [100, 123]). In this process, tumor
cells become less reliant on oxidative phosphorylation by
mitochondria and instead rely on conversion of pyruvate to
lactate by lactate dehydrogenase A and inactivation of pyru-
vate dehydrogenase, yielding just two molecules of ATP per
molecule of glucose [101, 123]. The heavy dependence on
aerobic glycolysis as the sole source of energy is thought to
arise as a survival mechanism during anaerobic conditions
within the hypoxic tumor; however, even in the presence of
oxygen tumor cells still turn to aerobic glycolysis. A theory
behind the preference is to maximize proliferation while also
providing the building blocks for the anabolic synthesis of
biomass in the form of nucleic acids, lipids and amino acids
[100, 101].
At a critical point in the decision towards aerobic glycolysis
is AMPK [22, 23]. Normally, AMPK is activated during
periods of low nutrient availability and inhibits TORC1 by
promoting the activity of the GTPase complex TSC1/TSC2,
effectively the opposite result of AKT activation and its inhib-
itory phosphorylation of TSC1/TSC2. This ensures that pro-
tein synthesis is reduced when insufficient nutrients are pres-
ent, and shunts the cell towards optimized ATP production via
oxidative phosphorylation and catabolic metabolism [23].
Low TORC1 signaling is not the condition that a tumor cell
prefers and, not surprisingly, the AMPK axis therefore func-
tions as a linear tumor suppressor pathway, by restricting
TORC1 signaling. For example, LKB1 is deleted or mutated
in sporadic cancer and is mutated in the familial Peutz-Jeghers
syndrome [124, 125]. Likewise, TSC1 and TSC2 are also
tumor suppressors and are mutated in the harmatoma syn-
drome Tuberous Sclerosis Complex [126]. Loss of phospha-
tase with pleckstrin homology on chromosome 10 (PTEN), a
genetic event that occurs in the majority of human cancers,
ameliorates the suppression of TORC1 by AMPK. Thus,
AMPK-activation of nutrient-dependent checkpoints is over-
come in some cancers as a means to bypass growth inhibition.
Importantly, induced activation of AMPK [22, 127]orre-
expression of LKB1 [128] in tumor cells reduces aerobic
glycolysis and slows tumor growth, indicating that AMPK
directly impacts on the decision to revert to glycolysis.
Mechanistically, repression of cell growth by adiponectin
could then involve a divergence of a tumor cell away from
glycolysis, thereby acting to slow cell growth and regulate
energy homeostasis under these conditions by repressing ini-
tiation of protein translation, elevating glucose uptake and
increasing fatty-acid oxidation [7, 24, 30, 129].
Future work will be needed to establish genetic models that
precisely activate or disrupt adiponectin/APPL1/AMPK sig-
naling in the context of tumor evolution that relies on aerobic
glycolysis. While the knockout of APPL1 has resulted in a
surprisingly mild phenotype under normal conditions [93], it
would be interesting to use this genetic background and others
such as AdipoR1/2 knockout to evaluate breast tumor growth
and metastasis where the long term down-regulation of
adiponectin signaling could be evaluated.
1.7 Adiponectin and a potential positive effect on breast
cancer
Although current evidence suggests that adiponectin sup-
presses tumor initiation and progression, there are some inter-
esting observations that might suggest t hat adiponectin c ould
also play a positive role. Takahata and colleagues reported that
Adiponectin
20 - 2 ug/ml
Direct
signaling
effects
Crosstalk
with insulin
signaling
Beneficial metabolic effects
Healthy individual
Defective adiponectin action
Metabolic dysfunction
Diabetes
CVD
Cancer
Reduced circulating or
local adiponectin levels
Adiponectin resistance
Fig. 2 Adiponectin action is altered in normal weight and obese individ-
uals. First, positive direct and insulin-sensitizing actions of adiponectin
are lost in obesity. One reason is that the normally high circulating
adiponectin concentration is reduced in obesity. Obesity may also induce
defective adiponectin action (adiponectin resistance) leading to metabolic
dysfunction. Ensuing disease states can include cardiovascular disease
(CVD), diabetes and cancer
Rev Endocr Metab Disord
AdipoR1 is present on breast cancer cell lines (MCF7, T47D,
MDA-MD231) and breast canc er tissue biopsies [130] breast
cancer tumors, and Karaduman studied 27 breast cancer pa-
tients and found that the localized concentrations of adiponectin
in breast tissue was elevated significantly compared to control
subjects [131]. Thus, compared with an overall trend towards
lower adiponectin serum levels in cancer patients, the localized
microenvironment could present a very different picture.
Specifically, we hypothesize that tumors, if they have ge-
netically inactivated the LKB1-AMPK axis, could benefit
from adiponectin stimulation by synergistically stimulating
insulin-and IGF1-induced signals; for example they could
experience increased PI3K/AKT activation and increased
Glut4 translocation and glucose uptake, and increased
recycling of insulin receptors to the plasma membrane via
Rab5. We performed meta-analysis of The Cancer Genome
Atlas (TCGA) that showed that AdipoR1 mRNA was
upregulated in 22 % of luminal A/B tumors (70 of 324).
Importantly, there was a significant reduction in survival time
in cases with upregulated AdipoR1 (P=0.004; Fig. 2). Clearly,
AdipoR1 protein expression, effects on downstream effectors,
and analysis of other genomic alterations are needed to under-
stand this preliminary observation. However, it does suggest
that adiponectin could play a role in the acceleration of some
types of luminal breast cancers, and argues that further inves-
tigation of this pathway is warranted.
In summary, it is clear from the above information that
correlations exist between adiponectin expression or changes
in adiponectin action and diseases such as cancer, cardiovas-
cular disease and diabetes. In particular, the adiponectin sig-
naling axis of AdipoR1 or AdipoR2, APPL1 and AMPK
appear to play a central role in many important actions of
adiponectin. Although current drugs can act at least in part via
enhancing adiponectin action, for example thiazolidinediones,
it is to be hoped that our current detailed understanding of
adiponectin signaling can be harnassed to develop new ther-
apeutics which can combat the clinical outcomes of disturbed
adiponectin function.
Acknowledgements Related work in the authors laboratories has been
supported by Canadian Institutes of Health Research, Canadian Diabetes
Association and Heart & Stroke Foundation of Canada. GS acknowl-
edges a Career Investigator Award from Heart & Stroke Foundation of
Ontario. We thank Hana Kim for graphic design assistance in Fig. 2.
Conflict of interest The authors declare that there is no conflict of
interest pertaining to the content of this review article.
References
1. Vucenik I, Stains JP. Obesity and cancer risk: Evidence, mecha-
nisms, and recommendations. [ Review]. Ann N Y Acad Sci.
2012;1271:3743. doi:10.1111/j.1749-6632.2012.06750.x.
2. Despres JP. Body fat distribution and risk of cardiovascular disease: An
update. [Research suppo rt, Non-U.S. Govt Revie w]. Circulation.
2012;126(10):130113. doi:10.1161/CIRCULATIONAHA.111.
067264.
3. Harwood Jr HJ. The adipocyte as an endocrine organ in the regu-
lation of metabolic homeostasis. Neuropharmacology.
2012;63(1):5775. doi:10.1016/j.neuropharm.2011.12.010.
4. Dadson K, Liu Y, Sweeney G. Adiponectin action: A combination
of endocrine and autocrine/paracrine effects. Front Endoc rinol
(Lausanne). 2011;2:62. doi:10.3389/fendo.2011.00062.
5. Hui X, Lam KS, Vanhoutte PM, Xu A. Adiponectin and cardiovas-
cular health: An update. Br J Pharmacol. 2012;165(3):57490.
doi:10.1111/j.1476-5381.2011.01395.x.
6. Dalamaga M, Diakopoul os K N, M antzoros CS. The role of
adiponectin in cancer: A review of current evidence. Endocr Rev.
2012;33(4):54794. doi:10.1210/er.2011-1015.
7. Yamauchi T, Kamon J, Ito Y, Tsuchida A, Yokomizo T, Kita S, et al.
Cloning of adiponectin receptors that mediate Antidiabetic meta-
bolic effects. [Research support, Non-U.S. Govt]. Nature.
2003;423(6941):7629. doi:10.1038/nature01705.
8. Hall J, Roberts R, Vora N. Energy homoeostasis: The roles of adipose
tissue-derived hormones, peptide YY and Ghrelin. Obesity Facts.
2009;2(2):11725.
9. Kadowaki T, Yamauchi T. Adiponectin and adiponectin receptors.
Endocr Rev. 2005;26(3):43951. doi:10.1210/er.2005-0005.
10. Jäger S, Handschin C, St.-Pierre J, Spiegelman BM. AMP-activated
protein kinase (AMPK) action in skeletal muscle via direct phos-
phorylation of PGC-1alpha. Proc Natl Acad Sci.
2007;104(29):1201722. doi:10.1073/pnas.0705070104.
11. Nemoto S, Fer gusson MM, Finkel T. SIRT1 Functionally interacts
with the metabolic regulator and transcriptional coactivator PGC-
1α. J Biol Chem. 2005;280(16):1645660. doi:10.1074/jbc.
M501485200.
12. Rodgers JT, Lerin C, Haas W, Gygi SP, Spiegelman BM, Puigserver
P. Nutrient control of glucose homeostasis through a complex of
PGC-1alpha and SIRT1. Nature. 2005;434(7029):1138. doi:10.
1038/nature03354.
13. Canto C, Gerhart-Hines Z, Feige JN, Lagouge M, Noriega L, Milne
JC, et al. AMPK regulates ene rgy expenditure by modulating
NAD+metabolism and SIRT1 activity. Nature. 2009;458(7241):
105660. doi:10.1038/nature07813.
14. Hardie DG, Alessi DR. LKB1 And AMPK and the cancer-metabolism
linkten years after . BMC Biol. 2013;1 1:36. doi:10.1186 /1741-7007-
11-36.
15. Zeqiraj E, Fi lippi BM, Deak M, Alessi DR, van Aalten DM.
Structure of the LKB1-STRAD-MO25 complex reveals an alloste-
ric me chanism of kinase activation. Sc ience. 2009;326(5960):
170711. doi:10.1126/science.1178377.
16. Boudeau J, Baas AF, Deak M, Morrice NA, Kieloch A,
Schutkowski M, et al. MO25alpha/beta interact with
STRADalpha/beta enhancing their ability to bind, activate and
localize LKB1 in the cytoplasm. EMBO J. 2003;22(19):510214.
doi:10.1093/emboj/cdg490.
17. Baas AF, Boudeau J, Sapkota GP, Smit L, Medema R, Morrice NA,
et al. Activation of the tumour suppressor kinase LKB1 by the
STE20-like pseudokinase STRAD. EMBO J. 2003;22(12):3062
72. doi:10.1093/emboj/cdg292.
18. Woods A, Dickerson K, Heath R, Hong SP, Momcilovic M,
Johnstone SR, et al. Ca2+/calmodulin-dependent protein kinase
kinase-beta acts upstream of AMP-activated protein kinase in mam-
malian cells. Cell Metab. 2005;2(1):2133. doi:10.1016/j.cmet.2005.
06.005.
19. Hawley SA, Pan DA, Mustard KJ, Ross L, Bain J, Edelman AM,
et al. Calmodulin-dependent protein kinase kinase-beta is an alter-
native upstream kinase for AMP-activated protein kinase. Cell
Metab. 2005;2(1):919. doi:10.1016/j.cmet.2005.05.009.
Rev Endocr Metab Disord
20. Herrero-Martin G, Hoyer-Hansen M, Garcia-Garcia C, Fumarola C,
Farkas T, Lopez-Rivas A, et al. TAK1 activates AMPK-dependent
cytoprotective autophagy in TRAIL-treated epithelial cells. EMBO
J. 2009;28(6):67785. doi:10.1038/emboj.2009.8.
21. Momcilovic M, Hong SP, Carlson M. Mammalian TAK1 activates
Snf1 protein kinase in yeast and phosphorylates AMP-activated
protein kinase in vitro. J Biol Chem. 2006;281(35):2533643.
doi:10.1074/jbc.M604399200.
22. Faubert B, Boily G, Izreig S, Griss T, Samborska B, Dong Z, et al.
AMPK is a negative regulator of the Warburg effect and suppresses
tumor growth in vivo. Cell Metab. 2013;17(1):11324.
23. Dandapan i M , Hardie DG. AMPK : Oppo sing t he m etabolic
changes in both tumour cells and inflammatory cells? Biochem
Soc Trans. 2013;41(2):68793. doi:10.1042/bst20120351.
24. Iwabu M, Yamauchi T, Okada-Iwabu M, Sato K, Nakagawa T,
Funata M, et al. A diponectin and AdipoR1 regulate PGC- 1alpha
and mitochondria by Ca(2+) and AMPK/SIR T1. [Research support,
Non-U.S. Govt]. Nature. 2010;464(7293):13139. doi:10.1038/
nature08991.
25. Wang C, Mao X, Wang L, Liu M, Wetzel MD, Guan KL, et al.
Adiponectin sensitizes insulin signaling by reducing p70 S6 kinase-
mediated serine phosphorylation of IRS-1. [Research Support,
N.I.H., Extramural Research Support, Non-U.S. Govt]. The
Journal of biological chemistry. 2007;282(11):79916. doi:10.
1074/jbc.M700098200.
26. Moore T, Beltran L, Carbajal S, Strom S, Traag J, Hursting SD, et al.
Dietary energy balance modulates signaling through the Akt/
mammalian target of rapamycin pathways in multiple epithelial
tissues. Cancer prevention research. 2008;1(1):6576. doi:10.
1158/1940-6207.capr-08-0022.
27. Taliaferro-Smith L, Nagalingam A, Zhong D, Zhou W, Saxena NK,
Sharma D. LKB1 is required for adiponectin-mediated modulation
of AMPK-S6K axis and inhibition of migration and invasion of
breast cancer cells. Oncogene. 2009;28(29):262133. doi:10.1038/
onc.2009.129.
28. Mao X, Kikani CK, Riojas RA, Langlais P, Wang L, Ramos FJ,
et a l. AP PL1 binds to adiponectin receptors and mediates
adiponectin signaling and function. [Research Support, N.I.H.,
Extramural Research Support, Non-U.S. Govt]. Nature cell biolo-
gy . 2006;8(5):51623. doi:10.1038/ncb1404.
29. Kubota N, Yano W, Kubota T, Yamauchi T, Itoh S, Kumagai H,
et al. Adiponectin stimulates AMP-activated protein kinase in the
hypothalamus and increases food intake. Cell Metab. 2007;6(1):55
68. doi:10.1016/j.cmet.2007.06.003.
30. Yamauchi T, Kamon J, Minokoshi Y, Ito Y, Waki H, Uchida S, et al.
Adiponectin stimulates glucose utilization and fatty-acid oxidation
by activating AMP-activated protein kinase. Nature medicine.
2002;8(11):128895. doi:10.1038/nm788.
31. Zhou L, Deepa SS, Etzler JC, Ryu J, Mao X, Fang Q, et al.
Adiponectin activates AMP-activated protein kinase in muscle cells
via APPL1/LKB1-dependent and Phospholipase C/Ca2+/Ca2+/cal-
modulin-dependent protein kinase kinase-dependent pathways. J Biol
Chem. 2009;284(33):22426
35 . doi:10.1074/jbc.M109 .028357.
32. Mitsuuchi Y, Johnson SW, Sonoda G, T anno S, Golemis EA, Testa
JR. Identification of a chromosome 3p14.3-21.1 gene, APPL,
encoding an adaptor molecule that interacts with the oncoprotein-
serine/threonine kinase AKT2. Oncogene. 1999;18(35):48918.
doi:10.1038/sj.onc.1203080.
33. Boudeau J, Scott JW, Resta N, Deak M, Kieloch A, Komander D,
et al. Analysis of the LKB1-STRAD-MO25 complex. J Cell Sci.
2004;117(Pt 26):636575. doi:10.1242/jcs.01571.
34. Xie Z, Dong Y, Zhang J, Scholz R, Neumann D, Zou MH.
Identification of the serine 307 of LKB1 as a novel phosphorylation
site essential for its nucleocytoplasmic transport and endothelial cell
angiogenesis. Mol Cell Biol. 2009;29(13):358296. doi:10.1 128/mcb.
01417-08.
35. Deepa SS, Zhou L, Ryu J, Wang C, Mao X, Li C, et al. APPL1
Mediates adiponectin-induced LKB1 cytosolic localization through
the PP2A-PKCΠsignaling pathway. Mol Endocrinol. 2011;25
(10):177385. doi:10.1210/me.2011-0082.
36. Manning BD, Cantley LC. AKT/PKB signaling: Navigating down-
stream. Cell. 2007;129(7):126174. doi:10.1016/j.cell.200 7.06.
009.
37. Scheid MP, Woodgett JR. Unravelling the activation mechanisms of
protein kinase B/Akt. FEBS Lett. 2003;546(1):10812.
38. Patel SA, Hoehn KL, Lawrence RT, Sawbridge L, Talbot NA,
Tomsig JL, et al. Overexpression of the adiponectin receptor
AdipoR1 in rat skeletal muscle amplifies local insulin sensitivity.
Endocrinology. 2012;153(11):52314 6. doi:10.1210/en.2012-
1368.
39. Maruyama S, Shibata R, Ohashi K, Ohashi T, Daida H, Walsh K,
et al. Adiponectin ameliorates doxorubicin-induced cardiotoxicity
through Akt protein-dependent mechanism. [Research Support,
Non-U.S. Govt]. The Journal of biological chemistry.
2011;286(37):32790800. doi:10.1074/jbc.M1 11.245985.
40. Saito T, Jones CC, Huang S, Czech MP, Pilch PF. The interaction of
Akt with APPL1 is required for insulin-stimulated Glut4 transloca-
tion. J Biol Chem. 2007;282(44):322 807 . doi:10.1074/jbc.
M704150200.
41. Cheng KKY, Iglesias MA, Lam KSL, Wang Y, Sweeney G, Zhu W,
et al. APPL1 Potentiates Insulin-Mediated Inhibition of Hepatic
Glucose Production and Alleviates Diabetes via Akt Activation in
Mice. Cell Metab. 2009;9(5):41727.
42. Wijesekara N, Krishnamurthy M, Bhattacharjee A, Suhail A,
Sweeney G, Wheeler MB. Adiponectin-induced ERK and Akt
phosphorylation protects against pancreatic beta cell apoptosis and
increases insulin gen e expression and secretion. J Biol Chem.
2010;285(44):3362331. doi:10.1074/jbc.M109.085084.
43. Tsuchida A, Yamauchi T, Ito Y, Hada Y, Maki T, Takekawa S, et al.
Insulin/Foxo1 pathway regulates expression levels of adiponectin
receptors and adiponectin sensitivity. J Biol Chem.
2004;279(29):3081722. doi:10.1074/jbc.M402367200.
44. Backer JM, Myers Jr MG, Shoelson SE, Chin DJ, Sun XJ, Miralpeix
M, et al. Phosphatidylinositol 3
-kinase is activated by association with
IRS-1 during insulin stimulation. EMBO J. 1992;11(9):346979.
45. Cornu M, Albert V, Hall MN. mTOR in aging, metabolism, and
cancer. Curr Opin Genet Dev. 2013;23(1):5362. doi:10.1016/j.gde.
2012.12.005.
46. Magnuson B, Ekim B, Fingar DC. Regulation and function of
ribosomal protein S6 kinase (S6K) within mTOR signaling net-
works. Biochem J. 2012;441(1):121. doi:10.1042/bj20110892.
47. Werner ED, Lee J, Hansen L, Yuan M, Shoelson SE. Insulin
resistance Due to phosphorylation of insulin receptor substrate-
1 at serine 302. J Biol Chem. 2004;279(34):35298305. doi:10.
1074/jbc.M405203200.
48. Harrington LS, Findlay GM, Gray A, Tolkacheva T, Wigfield S,
Rebholz H, et al. The TSC1-2 tumor suppressor controls insulin
ÄìPI3K signaling via regulation of IRS proteins. The Journal of
Cell Biology. 2004;166(2):21323. doi:10.1083/jcb.200403069.
49. Vu V, Kim W, Fang X, Liu YT, Xu A, Sweeney G. Coculture with
primary visceral rat adipocytes from control but not Streptozotocin-
induced diabetic animals increases glucose uptake in rat skeletal
muscle cells: Role of adiponectin. [Research support, Non-U.S.
Govt]. Endocrin ology. 2007;148(9):441 19. doi:10.1210/en.2007-
0020.
50. Gaddam KK, Ventura HO, Lavie CJ. Metabolic syndrome and heart
failurethe risk, paradox, and treatment. [Review]. Curr Hypertens
Rep. 2011;13(2):1428. doi:10.1007/s11906-011-0179-x.
51. Anker SD, von Haehling S. The obesity paradox in heart failure:
Accepting reality and making rational decisions. [Research support,
Non-U.S. Govt Review]. Clin Pharmacol Ther. 2011;90(1):188
90. doi:10.1038/clpt.2011.72.
Rev Endocr Metab Disord
52. Park M, Sweeney G. Direct effects of adipokines on the heart: Focus on
adiponectin. Heart Fail Rev . 2013. doi:10.1007/s10741-012-9337-8.
53. Fujita K, Maeda N, Sonoda M, Ohashi K, Hibuse T, Nishizawa H,
et al. Adiponectin protects against angiotensin II-induced cardiac
fibrosis through activation of PPAR-alpha. Arterioscler Thromb
Vasc Biol. 2008;28(5):86370.
54. Shibata R, Sato K, Pimentel DR, Takemura Y, Kihara S, Ohashi K,
et al. Adiponectin protects against myocardial ischemia-reperfusion
injury through AMPK- and COX-2-dependent mechanisms. Nat
Med. 2005;11(10):1096103.
55. Shibata R, Ouchi N, Ito M, Kihara S, Shiojima I, Pimentel DR, et al.
Adiponectin-mediated modulation of hypertrophic signals in the
heart. Nat Med. 2004;10(12):13849.
56. Shibata R, Izumiya Y, Sato K, Papanicolaou K, Kihara S, Colucci WS,
et al. Adiponectin protects against the development of systolic dysfunc-
tion following myocardial infarction. J Mol Cell Cardiol.
2007;42(6):106574.
57. Sam F, Duhaney TA, Sato K, Wilson RM, Ohashi K, Sono-
Romanelli S, et al. Adiponectin deficiency, diastolic dysfunction,
and diastolic heart failure. [Research Support, N.I.H., Extramural
Research Support, Non-U.S. Govt]. Endocrinology. 2010;151(1):
32231. doi:10.1210/en.2009-0806.
58. Shimano M, Ouchi N, Shibata R, Ohashi K, Pimentel DR,
Murohara T, et al. Adiponectin deficiency exacerbates cardiac
dysfunction following pressure overload through disruption of an
AMPK-dependent angiogenic response. [Research Support, N.I.H.,
Extramural Research Support, Non-U.S. Govt]. J Mol Cell
Cardiol. 2010;49(2):21020. doi:10.1016/j.yjmcc.2010.02.021.
59. Leroith D. Pathophysiology of the metabolic syndrome: implica-
tions for the cardiometabolic risks associated with type 2 diabetes.
[Review]. Am J Med Sci. 2012;343(1):136. doi:10.1097/MAJ.
0b013e31823ea214.
60. Abel ED, Litwin SE, Sweeney G. Cardiac remodeling in obesity.
Physiol Rev. 2008;88(2):389419.
61. Despres JP, Cartier A, Cote M, Arsenault BJ. The concept of
cardiometabolic risk: Bridging the fields of diabetology and cardi-
ology. [Research Support, Non-U.S. Govt Review]. Ann Med.
2008;40(7):51423. doi:10.1080/07853890802004959.
62. Fang X, Palanivel R, Cresser J, Schram K, Ganguly R, Thong FS,
et al. An APPL1-AMPK signaling axis mediates beneficial meta-
bolic effects of adiponectin in the heart. [In Vitro Research Support,
N.I.H., Extramural Research Support, Non-U.S. Govt]. Am J
Physiol Endocrinol Metab. 2010;299(5):E721729. doi:10.1152/
ajpendo.00086.2010.
63. Palanivel R, Fang X, Park M, Eguchi M, Pallan S, De Girolamo S,
et al. Globular and full-length forms of adiponectin mediate specific
changes in glucose and fatty acid uptake and metabolism in
cardiomyocytes. Cardiovasc Res. 2007;75(1):14857.
64.ParkM,YounB,ZhengXL,WuD,XuA,SweeneyG.Globular
adiponectin, acting via AdipoR1/APPL1, protects H9c2 cells from
hypoxia/reoxygenation-induced apoptosis. [Research Support, Non-
U.S. Govt]. PLoS One. 2011;6(4):e19143. doi:10.1371/journal.pone.
0019143.
65. Wang C, Xin X, Xiang R, Ramos FJ, Liu M, Lee HJ, et al. Yin-Yang
regulation of adiponectin signaling by APPL isoforms in muscle
cells. J Biol Chem. 2009;284 (46):3160815. doi:10.1074/jbc.
M109.010355.
66. Ebert T, Fasshauer M. Adiponectin: Sometimes good, sometimes
bad? Cardiology. 2011;118(4):2367. doi:10.1159/000329647.
67. Khan RS, Kato TS, Chokshi A, Chew M, Yu S, Wu C, et al.
Adipose tissue inflammation and adiponectin resistance in patients
with advanced heart failure: Correction after ventricular assist de-
vice implantation. Circ Heart Fail. 2012;5(3):3408. doi:10.1161/
CIRCHEARTFAILURE.111.964031.
68. Skurk C, Wittchen F, Suckau L, Witt H, Noutsias M, Fechner H,
et al. Description of a local cardiac adiponectin system and its
deregulation in dilated cardiomyopathy. [Research Support, Non-
U.S. Govt]. Eur Heart J. 2008;29(9):116880 . doi:10.1093/eurheartj/
ehn136.
69. Ma Y, Liu Y, Liu S, Qu Y, W ang R, Xia C, et al. Dynamic alteration of
adiponectin/adiponectin receptor expression and its impact on myocar-
dial ischemia/reperfusion in type 1 diabetic mice. Am J Physiol
Endocrinol Metab. 2011;301(3):E447455. doi:10.1152/ajpendo.
00687.2010.
70. Wang T, Qiao S, Lei S, Liu Y, Ng KF, Xu A, et al. N-acetylcysteine
and allopurinol synergistically enhance cardiac adiponectin content
and reduce myocardial reperfusion injury in diabetic rats. PLoS
One. 2011;6(8):e23967. doi:10.1371/journal.pone.0023967.
71. Cui XB, Wang C, Li L, Fan D, Zhou Y, Wu D, et al. Insulin
decreases myocardial adiponectin receptor 1 expression via PI3K/
Akt and FoxO1 pathway. [Research Support, Non-U.S. Govt].
Cardiovasc Res. 2012;93(1):6978. doi:10.1093/cvr/cvr273.
72. Fang X, Palanivel R, Zhou X, Liu Y, Xu A, Wang Y, et al.
Hyperglycemia-and hyperinsulinemia-induced alteration of
adiponectin receptor expression and adiponectin effects in L6 myo-
blasts. [Research Support, Non-U.S. Govt]. J Mol Endocrinol.
2005;35(3):46576. doi:10.1677/jme.1.01877.
73. Yi W, Sun Y, Gao E, Wei X, Lau WB, Zheng Q, et al. Reduced
cardioprotective action of adiponectin in high-fat diet-induced type
II diabetic mice and its underlying mechanisms. [Research Support,
N.I.H., Extramural Research Support, Non-U.S. Govt]. Antioxid
Redox Signal. 2011;15(7):177988. doi:10.1089/ars.2010.3722.
74. Wang C, Li L, Zhang ZG, Fan D, Zhu Y, Wu LL. Globular
adiponectin inhibits angiotensin II-induced nuclear factor kappaB
activation through AMP-activated protein kinase in cardiac hyper-
trophy. [Research Support, Non-U.S. Govt]. J Cell Physiol.
2010;222(1):14955. doi:10.1002/jcp.21931.
75. Cheng KK, Lam KS, Wang Y, Huang Y, Carling D, Wu D, et al.
Adiponectin-induced endothelial nitric oxide synthase activation
and nitric oxide production are mediated by APPL1 in endothelial
cells. Diabetes. 2007;56(5):138794. doi:10.2337/db06-1580.
76. Wu X, Mahadev K, Fuchsel L, Ouedraogo R, Xu SQ, Goldstein BJ.
Adiponectin suppresses IkappaB kinase activation induced by tu-
mor necrosis factor-alpha or high glucose in endothelial cells: role of
cAMP and AMP kinase signaling. Am J Physiol Endocrinol Metab.
2007;293(6):E18361844. doi:10.1152/ajpendo.00115.2007.
77. Ouchi N, Kihara S, Arita Y, Okamoto Y, Maeda K, Kuriyama H,
et al. Adiponectin, an adipocyte-derived plasma protein, inhibits
endothelial NF-kappaB signaling through a cAMP-dependent path-
way. Circulation. 2000;102(11):1296301.
78. Wang Y, Cheng KK, Lam KS, Wu D, Huang Y, Vanhoutte PM, et al.
APPL1 counteracts obesity-induced vascular insulin resistance and
endothelial dysfunction by modulating the endothelial production of
nitric oxide and endothelin-1 in mice. Diabetes. 2011;60(11):3044
54. doi:10.2337/db11-0666.
79. Schmid PM, Resch M, Steege A, Fredersdorf-Hahn S, Stoelcker B,
Birner C, et al. Globular and full-length adiponectin induce NO-
dependent vasodilation in resistance arteries of Zucker lean but not
Zucker diabetic fatty rats. Am J Hypertens. 2011;24(3):2707.
doi:10.1038/ajh.2010.239.
80. Turer AT, Scherer PE. Adiponectin: Mechanistic insights and clin-
ical implications. Diabetologia. 2012;55(9):231926. doi:10.1007/
s00125-012-2598-x.
81. Yamauchi T, Kadowaki T. Adiponectin receptor as a key player in
healthy longevity and obesity-related diseases. Cell Metab.
2013;17(2):18596. doi:10.1016/j.cmet.2013.01.001.
82. Hug C, Wang J, Ahmad NS, Bogan JS, Tsao TS, Lodish HF. T-
cadherin is a receptor for hexameric and high-molecular-weight
forms of Acrp30/adiponectin. Proc Natl Acad Sci U S A.
2004;101(28):1030813. doi:10.1073/pnas.0403382101.
83. Ordelheide, A. M., Heni, M., Gommer, N., Gasse, L., Haas, C.,
Guirguis, A., et al. The myocyte expression of adiponectin receptors
Rev Endocr Metab Disord
and PPARdelta is highly coordinated and reflects lipid metabolism
of the human donors. [Research Support, Non-U.S. Govt]. Exp
Diabetes Res, 2011, 692536, doi:10.1155/2011/692536.
84. Ceddia RB, Somwar R, Maida A, Fang X, Bikopoulos G, Sweeney
G. Globular adiponectin increases GLUT4 translocation and glu-
cose uptake but reduces glycogen synthesis in rat skeletal muscle
cells. Diabetologia. 2005;48(1):1329. doi:10.1007/s00125-004-
1609-y.
85. Liu Y, Chewchuk S, Lavigne C, Brule S, Pilon G, Houde V, et al.
Functional significance of skeletal muscle adiponectin production,
changes in animal models of obesity and diabetes, and regulation by
rosiglitazone treatment. Am J Physiol Endocrinol Metab.
2009;297(3):E657664. doi:10.1152/ajpendo.00186.2009.
86. Wang Y, Zhou M, Lam KS, Xu A. Protective roles of adiponectin in
obesity-related fatty liver diseases: Mechanisms and therapeutic
implications. Arq Bras Endocrinol Metabol. 2009;53(2):20112.
87. Ma K, Cabrero A, Saha PK, Kojima H, Li L, Chang BH, et al.
Increased beta -oxidation but no insulin resistance or glucose intol-
erance in mice lacking adiponectin. J Biol Chem. 2002;277(38):
3465861. doi:10.1074/jbc.C200362200.
88. Maeda N, Shimomura I, Kishida K, Nishizawa H, Matsuda M,
Nagaretani H, et al. Diet-induced insulin resistance in mice lacking
adiponectin/ACRP30. Nature medicine. 2002;8(7):7317. doi:10.
1038/nm724.
89. Kubota N, Terauchi Y, Yamauchi T, Kubota T, Moroi M, Matsui J,
et al. Disruption of adiponectin causes insulin resistance and
neointimal formation. J B iol Chem. 2002;277(29):258636.
doi:10.1074/jbc.C200251200.
90. Yamauchi T, Nio Y, Maki T, Kobayashi M, Takazawa T, Iwabu M,
et al. Targeted disruption of AdipoR1 and AdipoR2 causes abroga-
tion of adiponectin binding and metabolic actions. Nat Med.
2007;13(3):3329. doi:10.1038/nm1557.
91. Bjursell M, Ahnmark A, Bohlooly YM, William-Olsson L, Rhedin
M, Peng XR, et al. Opposing effects of adiponectin receptors 1 and
2 on energy metabolism. Diabetes. 2007;56(3):58393. doi:10.
2337/db06-1432.
92. Cleasby ME, Lau Q, Polkinghorne E, Patel SA, Leslie SJ, Turner N,
et al. The adaptor protein APPL1 increases glycogen accumulation
in rat skeletal muscle through activation of the PI3-kinase signaling
pathway. J Endocrinol. 2011. doi:10.1530/JOE-11-0039.
93. Tan Y, You H, Wu C, Altomare DA, Testa JR. Appl1 is dispensable
for mouse development, and loss of Appl1 has growth factor-
selective effects on Akt signaling in murine embryonic fibroblasts.
J Biol Chem. 2010;285(9):637789. doi:10.1074/jbc.M109.
068452.
94. Qiao L, Kinney B, Yoo HS, Lee B, Schaack J, Shao J. Adiponectin
increases skeletal muscle mitochondrial biogenesis by suppressing
mitogen-activated protein kinase phosphatase-1. [Research Support,
N.I.H., Extramural Research Support, Non-U.S. Govt]. Diabetes.
2012;61(6):146370. doi:10.2337/db1 1-1475.
95. Civitarese AE, Ukropcova B, Carling S, Hulver M, DeFronzo RA,
Mandarino L, et al. Role of adiponectin in human skeletal muscle
bioenergetics. Cell Metab. 2006;4(1):7587. doi:10.1016/j.cmet.
2006.05.002.
96. Vu V, R iddell MC , Sweeney G. Circulating adiponectin and
adiponectin receptor expression in skeletal muscle: Effects of exer-
cise. Diabetes Metab Res Rev. 2007;23(8):600
11. doi:10.1002/
dmrr.778.
97. Marinho R, Ropelle ER, Cintra DE, De Souza CT, Da Silva AS,
Bertoli FC, et al. Endurance exercise training increases APPL1
expression and improves insulin signaling in the hepatic tissue of
diet-induced obese mice, independently of weight loss. [Research
Support, Non-U.S. Govt]. J Cell Physiol. 2012;227(7):291726.
doi:10.1002/jcp.23037.
98. Farias JM, Maggi RM, Tromm CB, Silva LA, Luciano TF, Marques
SO, et al. Exercise training performed simultaneously to a high-fat
diet reduces the degree of insulin resistance and improves adipoR1-
2/APPL1 protein levels in mice. [Research Support, Non-U.S.
Govt]. Lipids Health Dis. 2012;11:134. doi:10.1186/1476-511X-
11-134.
99. Ho lmes RM, Yi Z, De Filippis E, Berria R, Shahani S,
Sathyanarayana P, et al. Increased abundance of the adaptor protein
containing pleckstrin homology domain, phosphotyrosine binding
domain and leucine zipper motif (APPL1) in patients with obesity
and type 2 diabetes: evidence for altered adiponectin signaling.
Diabetologia. 2011. doi:10.1007/s00125-011-2173-x.
100. Koppenol WH, Bounds PL, Dang CV. Otto Warburgs contributions
to current concepts of cancer metabolism. Nat Rev Cancer.
2011;11(5):32537. doi:10.1038/nrc3038.
101. Levine AJ, Puzio-Kuter AM. The control of the metabolic switch in
cancers by on cogenes and tumor supp ressor genes. Scien ce.
2010;330(6009):13404. doi:10.1126/science.1193494.
102. VanSaun MN. Molecular pathways: Adiponectin and leptin signal-
ing in cancer. Clin Cancer Res. 2013;19(8):192632. doi:10.1158/
1078-0432.ccr-12-0930.
103. Bergstrom A, Pisani P, Tenet V, Wolk A, Adami HO. Overweight as
an avoidable cause of cancer in Europe. International journal of
cancer. Journal international du cancer. 2001;91(3):42130.
104. Kritchevsky D. Diet and cancer: Whats Next? The Journal of
nutrition. 2003;133(11):3827S9S.
105. Shehzad A, Iqbal W, Shehzad O, Lee YS. Adiponectin: Regulation
of its production and its role in human diseases. Hormones (Athens).
2012;11(1):820.
106. Hu E, Liang P, Spiegelman BM. AdipoQ is a novel adipose-specific
gene dysregulated in obesity. J Biol Chem. 1996;271(18):10697
703. doi:10.1074/jbc.271.18.10697.
107. Hotta K, Funahashi T, Bodkin NL, Ortmeyer HK, Arita Y, Hansen
BC, et al. Circulating concentrations of the adipocyte protein
adiponectin Are decreased in parallel with reduced insulin sensitiv-
ity during the progression to type 2 diabetes in rhesus monkeys.
Diabetes. 2001;50(5):112633. doi:10.2337/diabetes.50.5.1126.
108. Korner A, Pazaitou-Panayiotou K, Kelesidis T , Kelesidis I, Williams
CJ, Kaprara A, et al. Total and high-molecular-weight adiponectin in
breast cancer: In vitro and in vivo studies. Journal of Clinical
Endocrinology & Metabolism. 2007;92(3):10418. doi:10.1210/jc.
2006-1858.
109. Mantzoros C, Petridou E, Dessypris N, Chavelas C, Dalamaga M,
Alexe DM, et al. Adiponectin and breast cancer risk. Journal of
Clinical Endocrinology & Metabolism. 2004;89(3):1102
7. doi:10.
1210/jc.2003-031804.
110. Miyoshi Y, Funahashi T, Kihara S, Taguchi T, Tamaki Y,
Matsuzawa Y, et al. Association of serum adiponectin levels with
breast cancer risk. Clin Cancer Res. 2003;9(15):5699704.
111. Chen D-C, Chung Y-F, Yeh Y-T, Chaung H-C, Kuo F-C, Fu O-Y,
et al. Serum adiponectin and leptin levels in Taiwanese breast cancer
patients. Cancer letters. 2006;237(1):10914.
112. Duggan C, Irwin ML, Xiao L, Henderson KD, Smith AW,
Baumgartner RN, et al. Associations of insulin resistance and
adiponectin with mortality in women with breast cancer. J Clin
Oncol. 2011;29(1):329. doi:10.1200/jco.2009.26.4473.
113. Dieudonne M-N, Bussiere M, Dos Santos E, Leneveu M-C, Giudicelli
Y, Pecquery R. Adiponectin mediates antiproliferative and apoptotic
responses in human MCF7 breast cancer cells. Biochem Biophys Res
Commun. 2006;345(1):2719. doi:10.1016/j.bbrc.2006.04.076.
114. Grossmann ME, Nkhata KJ, Mizuno NK, Ray A, Cleary MP.
Effects of adiponectin on breast cancer cell growth and signaling.
Br J Cancer. 2008;98(2):3709. doi:10.1038/sj.bjc.6604166.
115. Li G, Cong L, Gasser J, Zhao J, Chen K, Li F. Mechanisms
underlying the anti-proliferative actions of adiponectin in human
breast cancer cells, MCF7ÄìDependency on the cAMP/protein
kinase-a pathway. Nutrition and Cancer. 2010;63(1):808. doi:10.
1080/01635581.2010.516472.
Rev Endocr Metab Disord
116. Pfeiler GH, Buechler C, Neumeier M, Schaffler A, Schmitz G,
Ortmann O, et al. Adiponectin effects on human breast cancer cells
are dependent on 17-beta estradiol. Oncol Rep. 2008;19(3):78793.
117. Taliaferro-Smith L, Nagalingam A, Knight BB, Oberlick E, Saxena
NK, Sharma D. Integral role of PTP1B in adiponectin-mediated
inhibition of oncogenic actions of leptin in breast carcinogenesis.
Neoplasia. 2013;15(1):2338.
118. Nkhata KJ, Ray A, Schuster TF, Grossmann ME, Cleary MP.
Effects of adiponectin and leptin co-treatment on human breast
cancer cell growth. Oncol Rep. 2009;21(6):16119.
119. Kim KY, Baek A, Hwang JE, Choi YA, Jeong J, Lee MS, et al.
Adiponect in-ac tivated AM PK st imulates dephosphorylation o f
AKT through protein phosphatase 2A activation. Cancer Res.
2009;69(9):401826. doi:10.1158/0008-5472.can-08-2641.
120. Wang Y, Lam J B, Lam KS, Liu J, Lam MC, Ho o RL, et al.
Adiponectin modulates the glycogen synthase kinase-3beta/beta-
catenin signaling pathway and attenuates mammary tumorigenesis
of MDA-MB-231 cells in nude mice. Cancer Res. 2006;66
(23):1146270. doi:10.1158/0008-5472.can-06-1969.
121. Liu J, Lam JB, Chow KH, Xu A, Lam KS, Moon RT, et al.
Adiponectin stimulates Wnt inhibitory factor-1 expression through
epigenetic regulations involving the transcription factor specificity
protein 1. Carcinogenesis. 2008;29(11):2195202. doi:10.1093/
carcin/bgn194.
122. Lam JB, Chow KH, Xu A, Lam KS, Liu J, Wong NS, et al.
Adiponectin haploinsufficiency promotes mammary tumor devel-
opment in MMTV-PyVT mice by modulation of phosphatase and
tensin homolog activities. PLoS One. 2009;4(3):e4968. doi:10.
1371/journal.pone.0004968.
123. Vander Heiden MG, Cantley LC, Thompson CB. Understanding
the Warburg effect: The metabolic requirements of cell prolifera-
tion. Science. 2009;324(5930):102933. doi:10.1126/science.
1160809.
124. Boudeau JRM, Sapkota G, Alessi DR. LKB1, a protein kinase
regulating cell proliferation and polarity. FEBS Lett. 2003;
546(1):15965. doi:10.1016/S0014-5793(03)00642-2.
125. Kyriakis JM. At the crossroads: AMP-activated kinase and the
LKB1 tumor suppressor link cell proliferation to metabolic regula-
tion. J Biol. 2003;2(4):26. doi:10.1186/1475-4924-2-26.
126. Inoki K, Zhu T, Guan K-L. TSC2 Mediates cellular energy response
to control cell growth and survival. Cell. 2003;115(5):57790.
127. Kimura N, Tokunaga C, Dalal S, Richardson C, Yoshino K, Hara K,
et al. A possible linkage between AMP-activated protein kinase
(AMPK) and mammalian target of rapamycin (mTOR) signaling
pathway. Genes Cells. 2003;8(1):6579.
128. Shaw RJ, Bardeesy N, Manning BD, Lopez L, Kos matka M,
DePinho RA, et al. The LKB1 tumor suppressor negatively regu-
lates mTOR signaling. Cancer cell. 2004;6(1):919.
129. Hardie DG. AMP-activated/SNF1 protein kinases: Conserved
guardians of cellular energy. Nat Rev Mol Cell Biol. 2007;8
(10):77485. doi:10.1038/nrm2249.
130. Takahata C, Miyoshi Y, Irahara N, Taguchi T, Tamaki Y, Noguchi S.
Demonstration of adiponectin receptors 1 and 2 mRNA expression
in human breast cancer cells. Cancer Lett. 2007;250(2):22936.
doi:10.1016/j.canlet.2006.10.006.
131. Karaduman M, Bilici A, Ozet A, Sengul A, Musabak U, Alomeroglu
M. Tissue levels of adiponectin in breast cancer patients. Med Oncol.
2007;24(4):3616.
Rev Endocr Metab Disord
... 16 A diferencia de la leptina, la adiponectina se caracteriza por sus efectos antiinflamatorios, antiaterogénicos e insulino-sensibilizantes. Sus niveles plasmáticos se encuentran regulados Wanionok N.E., et al.: Síndrome metabólico, metformina y hueso negativamente por la acumulación de grasa visceral y son menores en personas obesas y con SM. 25,26 De hecho, la relación adiponectina/ leptina se utiliza como marcador de disfunción del tejido adiposo, se correlaciona estrechamente con la resistencia a la insulina y es más eficaz que la medición de cada hormona por separado. Por lo tanto, esta relación se ha propuesto como un marcador predictivo del SM. 23 Inflamación crónica La resistencia a la insulina y el estrés oxidativo, caracterizados por estar presentes en el SM, desencadenan un estado proinflamatorio activando diferentes cascadas de señalización, que pueden evaluarse a través del incremento de distintos marcadores inflamatorios, entre los que encontramos: TNF-α, IL-6 y proteína C-reactiva (PCR). ...
Article
Full-text available
Resumen El síndrome metabólico se define como un trastorno heterogéneo y multifactorial con riesgo cardiovascular elevado. Actualmente se encuentra en franco crecimiento debido al sedentarismo y la ingesta rica en grasas y azúcares. Su tratamiento incluye la indicación de cambios en el estilo de vida, con realiza-ción de actividad física y una alimentación saludable e hipocalórica. Cuando esto no es eficaz, se pueden utilizar diferentes fármacos, y entre los más prescriptos se encuentra la metformina, caracterizada por su acción insu-lino-sensibilizante. Numerosos trabajos han estudiado la vincu-lación del síndrome metabólico con el tejido óseo. Se demostró como resultado general, aunque no concluyente, que dicho síndrome se asocia con una disminución de la densidad mineral ósea y un aumento en la incidencia de fracturas osteoporóticas. Una de las limitacio-nes de estos estudios clínicos estaría ligada a la gran heterogeneidad de los pacientes con síndrome metabólico. Por otra parte, y dado que diversos estudios preclínicos han suge-rido posibles acciones osteogénicas de la metformina, se ha investigado el posible efec-to óseo de un tratamiento con este fármaco en personas con hiperglucemia o disgluce-mia. Varios estudios clínicos muestran que este efecto sería nulo o, en algunos casos, de carácter protector para el sistema óseo. No obstante, se debería tener precaución en el uso de dicho fármaco en pacientes que ne-cesiten dosis altas y/o posean riesgo elevado de fractura, ya que sus altas concentraciones podrían tener consecuencias negativas sobre el metabolismo óseo. Palabras clave: densidad mineral ósea, frac-turas osteoporóticas, metformina, síndrome metabólico.
... Indeed, adiponectin mediates several functions in vascular endothelium, e.g., production of NO via phosphorylation of eNOS at Ser1177 or attenuation of production of ROS [18,19], suggesting that reduced levels of adiponectin might also be related to NO levels in aorta. Thus, we studied the expression of proteins related to the proper function of vascular endothelium. ...
Article
Full-text available
Background C-reactive protein (CRP) is an acute inflammatory protein detected in obese patients with metabolic syndrome. Moreover, increased CRP levels have been linked with atherosclerotic disease, congestive heart failure, and ischemic heart disease, suggesting that it is not only a biomarker but also plays an active role in the pathophysiology of cardiovascular diseases. Since endothelial dysfunction plays an essential role in various cardiovascular pathologies and is characterized by increased expression of cell adhesion molecules and inflammatory markers, we aimed to detect specific markers of endothelial dysfunction, inflammation, and oxidative stress in spontaneously hypertensive rats (SHR) expressing human CRP. This model is genetically predisposed to the development of the metabolic syndrome. Methods Transgenic SHR male rats (SHR-CRP) and non-transgenic SHR (SHR) at the age of 8 months were used. Metabolic profile (including serum and tissue triglyceride (TAG), serum insulin concentrations, insulin-stimulated incorporation of glucose, and serum non-esterified fatty acids (NEFA) levels) was measured. In addition, human serum CRP, MCP-1 (monocyte chemoattractant protein-1), and adiponectin were evaluated by means of ELISA, histological analysis was used to study morphological changes in the aorta, and western blot analysis of aortic tissue was performed to detect expression of endothelial, inflammatory, and oxidative stress markers. Results The presence of human CRP was associated with significantly decreased insulin-stimulated glycogenesis in skeletal muscle, increased muscle and hepatic accumulation of TAG and decreased plasmatic cGMP concentrations, reduced adiponectin levels, and increased monocyte chemoattractant protein-1 (MCP-1) levels in the blood, suggesting pro-inflammatory and presence of multiple features of metabolic syndrome in SHR-CRP animals. Histological analysis of aortic sections did not reveal any visible morphological changes in animals from both SHR and SHR-CRP rats. Western blot analysis of the expression of proteins related to the proper function of endothelium demonstrated significant differences in the expression of p-eNOS/eNOS in the aorta, although endoglin (ENG) protein expression remained unaffected. In addition, the presence of human CRP in SHR in this study did not affect the expression of inflammatory markers, namely p-NFkB, P-selectin, and COX2 in the aorta. On the other hand, biomarkers related to oxidative stress, such as HO-1 and SOD3, were significantly changed, indicating the induction of oxidative stress. Conclusions Our findings demonstrate that CRP alone cannot fully induce the expression of endothelial dysfunction biomarkers, suggesting other risk factors of cardiovascular disorders are necessary to be involved to induce endothelial dysfunction with CRP.
... The concentration of circulating leptin in the blood is high in obese patients, and chronic leptin secretion caused by obesity can induce leptin resistance [45]. Adiponectin is mostly expressed in adipose tissue, and plasma adiponectin concentrations decrease in patients with obesity or cardiovascular disease [46,47]. Studies have shown that adiponectin increases insulin sensitivity and stimulates fatty acid oxidation by preventing lipid accumulation in skeletal muscles and the liver. ...
Article
Full-text available
Obesity is one of the major risk factors for metabolic diseases worldwide. This study examined the effects of YC-1102, an extract derived from the roots of Rosa multiflora, on 3T3-L1 preadipocytes and high-fat diet (HFD)-induced obese mice. In vivo experiments involved the oral administration of YC-1102 (100, 150, and 200 mg/kg body weight) daily to mice for eight weeks. YC-1102 was found to downregulate the expressions of PPARγ and C/EBPα during adipogenesis, inhibiting adipocyte differentiation and upregulating the expression of PGC-1α for energy metabolism to enhance mitochondrial biogenesis and fatty acid oxidation. It has been shown that daily administration of YC-1102 to mice receiving a HFD prevented an increase in body weight and the accumulation of body fat. YC-1102 administration also reduced TG, TC, and LDL cholesterol levels, as well as glucose and leptin levels, and increased adiponectin levels, thus effectively inhibiting the metabolism of lipids. YC-1102-treated mice showed significant reductions in the mRNA expression of PPARγ and C/EBPα. The levels of PGC-1α involved in energy metabolism increased significantly in the YC-1102-treated mice when compared to the HFD-treated mice. According to the findings of this study, YC-1102 has a dual mechanism that reduces transcription factors that promote the differentiation of adipocytes and increases transcription factors that promote energy consumption.
... This is in line with previous research that found that rats treated with an HF diet for 24 and 32 weeks experienced increases in Adiponectin plasma levels (62,63). Adiponectin is known to have an insulinsensitizing effect, but obesity has been associated with a malfunction in Adiponectin signaling, or Adiponectin resistance (64). Our current study also showed that the use of tocotrienols reduced the levels of leptin and increased adiponectin, but was not significant in comparison with the diet-induced obese group, and this was consistent with what Kok-Yong found: a study showed that, in male rats receiving buserelin for a 12-week period of therapy, oral administration of annatto tocotrienol at 60 or 100 mg/kg had no effect on levels of adiponectin or leptin (65).Although another investigation found that delta tocotrienol reduced leptin protein content in the tocotrienol group at dose T400 whereas not in the tocotrienol group at dose T1600 groups when compared to high-fat-fed mice, there were no differences in serum levels of the anti-inflammatory adipokine adiponectin between any of the groups (22). ...
... Adiponectin plays a crucial role in regulating insulin sensitivity, lipid metabolism, and glucose levels and exerts anti-inflammatory, anti-fibrotic, and antioxidant properties via the activation of AMP-activated protein kinase (AMPK) and peroxisome proliferator-activated receptor alpha (PPARa) pathways, reducing reactive oxygen species, and increasing the activity of antioxidant enzymes [51,53,54]. Additionally, adiponectin demonstrates angiogenic and vasodilatory functions, contributing to vascular health [48]. ...
Article
Full-text available
Adipokines are protein hormones secreted by adipose tissue in response to disruptions in physiological homeostasis within the body’s systems. The regulatory functions of adipokines within the central nervous system (CNS) are multifaceted and intricate, and they have been identified in a number of pathologies. Therefore, specific adipokines have the potential to be used as biomarkers for screening purposes in neurological dysfunctions. The systematic review presented herein focuses on the analysis of the functions of various adipokines in the pathogenesis of CNS diseases. Thirteen proteins were selected for analysis through scientific databases. It was found that these proteins can be identified within the cerebrospinal fluid either by their ability to modify their molecular complex and cross the blood–brain barrier or by being endogenously produced within the CNS itself. As a result, this can correlate with their measurability during pathological processes, including Alzheimer’s disease, amyotrophic lateral sclerosis, multiple sclerosis, depression, or brain tumors.
... Liver hypertrophy and elevated plasma lipid and leptin levels observed in fructose-fed rats may cause a reduction in adiponectin production, decreasing the adiponectin/ leptin ratio and adiponectin secretion that would result in hepatic lipid accumulation, confirming the metabolic abnormalities caused by fructose. Similar results were observed in Korean adults with type 2 diabetes (Lee et al., 2009) and in MS patients (Scheid and Sweeney, 2014). ...
Article
Full-text available
This study evaluates the potential effects of pumpkin seeds protein on blood pressure (BP), plasma adiponectin, leptin levels, and oxidative stress in rats with fructose-induced metabolic syndrome. Twenty four male Wistar albino rats were divided into four groups and fed a 20% casein diet, 20% casein diet supplemented with pumpkin protein, 20% casein diet with 64% D-fructose, or 20% casein diet with pumpkin protein and 64% D-fructose for 8 weeks. Contin-uous fructose feeding induced an increase in plasma insulin/glucose ratio, BP, insulin and glucose, aspartate aminotrans-ferase, alanine aminotransferase (ALT), alkaline phosphatase (ALP), creatinine, urea, and uric acid levels, and a decrease in the liver and muscle glycogen stores. In addition, elevated levels of total cholesterol (TC), triglycerides (TG), and leptin and lowered adiponectin levels were observed in rats fed a fructose-enriched diet. These groups also exhibited lower plasma levels of ascorbic acid and glutathione, higher thiobarbituric acid-reactive substances, hydroperoxide, carbonyl, and nitric oxide in both the liver and kidneys than rats fed the control diet. Interestingly, pumpkin seed protein treatment significantly counteracted alterations induced by fructose improving glucose, insulin, BP, TG, TC, ALT, and ALP levels, increasing liver and muscle glycogen stores, adiponectin level, and adiponectin/leptin ratio, and reducing plasma leptin lev-els. In addition, rats fed pumpkin protein with a high-fructose diet improved oxidative stress in the liver and kidneys. In conclusion, proteins from Cucurbita pepo L. seeds effectively improve metabolic parameters and protect against oxidative stress induced by a high-fructose diet.
... Body adipose tissue has the biological functions of altering lipid metabolism, regulating fat factors and causing chronic inflammation. An increase in visceral adipose tissue (VAT) leads to an increase in proinflammatory cytokines, 13 a decrease in circulating adiponectin level, 14 and increases in leptin level 15 and the severity of insulin resistance. 16 The role of visceral fat in the development of RCC is more significant than that of subcutaneous fat. ...
Article
Full-text available
Purpose To overcome the challenge of preoperative differentiation between clear cell renal cell carcinoma (ccRCC) and renal angiomyolipoma with minimal fat (RMFAML), we evaluated the potential of visceral adipose tissue (VAT) in distinguishing RMFAML from ccRCC. Patients and Methods Patients (191) were divided into ccRCC and RMFAML groups according to postoperative pathology. Umbilical horizontal computed tomography (CT) images were used for visceral fat area (VFA), subcutaneous fat area (SFA) and total fat area (TFA) measurements. Logistic regression was used to identify risk factors for ccRCC. Areas under the receiver operating characteristic (ROC) curve (AUCs) were compared to identify the most valuable indicator for identifying ccRCC and RMFAML. Results In total, 166 patients had ccRCC, and 25 had RMFAML. ccRCC and RMFAML patients showed significant differences in age (P<0.001), sex (P<0.001), hypertension (P=0.027), BMI (P<0.001), SFA (P=0.046), VFA (P<0.001) and TFA (P<0.001). According to multiple logistic regression analysis, male sex [4.311 (1.469~12.653), p=0.008]; older age [1.047 (1.008~1.088), p=0.017]; and higher BMI [1.305 (1.088~1.566), p=0.004], SFA [1.013 (1.003~1.023), p=0.008], VFA [1.026 (1.012~1.041), p<0.001] and TFA [1.011 (1.005~1.017), p=0.001] were associated with ccRCC. The AUCs of sex (male), age, BMI, TFA, VFA, and SFA were 0.726, 0.687, 0.783, 0.769, 0.840, and 0.645, respectively. The VFA cut-off value was 69.99 cm². The sensitivity and specificity of higher VFA (≥69.99 cm²) for ccRCC diagnosis were 79.52% and 80.00%, respectively. Conclusion In differentiating ccRCC from RMFAML, male sex, older age, and higher BMI, TFA, SFA, and VFA are risk factors for ccRCC. VFA is the most effective indicator for identifying ccRCC.
... Leptin and adiponectin are adipokines produced by adipose tissue, with leptin being secreted in proportion to body fat involved in the regulation of several important physiological functions [57]. However, it has been revealed that plasma adiponectin concentrations are low in patients with obesity and also MetS [58]. Therefore, the adiponectin/leptin ratio has been proposed as a predictive marker for MS, as this ratio declines as the number of metabolic risk variables increases, reflecting adipose tissue functionality [59]. ...
Article
Introduction The prevalence of metabolic syndrome (MetS) and abdominal obesity is escalating in South Asian countries. It is well established that MetS is associated with increased risk for both Type 2 diabetes mellitus and cardiovascular diseases. South Asians have an increased risk of MetS due to a variety of factors including unhealthy lifestyle and their unique body composition. Areas covered In this review we discuss the prevalence, associated risk factors and evidence-based preventive and curative strategies for MetS and abdominal obesity in South Asians. A literature search through PubMed®, Web of Science® and Scopus® was performed for studies published before 31st April 2021. A combination of the following keywords was used with the names of the individual South Asian countries: “metabolic syndrome”, “syndrome X”, “abdominal obesity”, “central obesity”, “visceral obesity” “prevention” and “management”. Expert opinion According to current evidence MetS and abdominal obesity are highly prevalent among South Asians. Several risk factors such as lifestyle, socio-demography, cultural, and body composition are associated with MetS. A limited research shows culturally tailored lifestyle interventions are effective in preventing and managing MetS and abdominal obesity among South Asians.
... Notably, a recent study has shown that adiponectin inhibits the expression of GLUT-1 in placental cells [65]. Regardless of how glucose import is affected by adiponectin, current evidence appears to support the concept that adiponectin stimulates oxidative phosphorylation of glucose, but represses glycolysis and lactate production, which is opposite to the Warburg effect [65][66][67]. To date, many of the biological actions of adiponectin on metabolism have been shown to be mediated via AMPK activation [68]. In addition, phosphorylation of AMPK by adiponectin was demonstrated to be a critical event underlying its anti-cancer effects [69]. ...
Article
Full-text available
Adiposity is associated with an increased risk of various types of carcinoma. One of the plausible mechanisms underlying the tumor-promoting role of obesity is an aberrant secretion of adipokines, a group of hormones secreted from adipose tissue, which have exhibited both oncogenic and tumor-suppressing properties in an adipokine type- and context-dependent manner. Increasing evidence has indicated that these adipose tissue-derived hormones differentially modulate cancer cell-specific metabolism. Some adipokines, such as leptin, resistin, and visfatin, which are overproduced in obesity and widely implicated in different stages of cancer, promote cellular glucose and lipid metabolism. Conversely, adiponectin, an adipokine possessing potent anti-tumor activities, is linked to a more favorable metabolic phenotype. Adipokines may also play a pivotal role under the reciprocal regulation of metabolic rewiring of cancer cells in tumor microenvironment. Given the fact that metabolic reprogramming is one of the major hallmarks of cancer, understanding the modulatory effects of adipokines on alterations in cancer cell metabolism would provide insight into the crosstalk between obesity, adipokines, and tumorigenesis. In this review, we summarize recent insights into putative roles of adipokines as mediators of cellular metabolic rewiring in obesity-associated tumors, which plays a crucial role in determining the fate of tumor cells.
... Abdominal adiposity can vary significantly within a small range of BMI and total body fat. Excessive visceral adipose tissue is associated with decreased adiponectin levels, elevated levels of cytokines (IL-6, TNF-α), insulin resistance, systemic inflammation (1). In 2015, there were an estimated 603 million obese adults and 107 million obese children worldwide (2). ...
Article
For the past several decades, we have witnessed the emergence of the obesity pandemic worldwide and, simultaneously, the increase of incidence of malignant diseases. The effects of obesity and overweight on cancer incidence, morbidity, and mortality started to be meticulously researched only recently. According to the epidemiological data analysis, the connection between obesity and increased risk of numerous cancers has been established. Estimations are that a change in lifestyle and diet can prevent 30-50% of malignant diseases. After smoking, obesity is the second largest preventable cause of cancer. Obesity affects the quality of life and increases the risk of cancer recurrence and cancer-related mortality. By reducing body mass and avoiding gaining weight during adulthood, the risk of getting cancer is lowered. Numerous studies have shown the beneficial effects of physical activity during and after cancer treatment. Obesity influences cancer development; however, the mechanisms responsible for it are still unclear. It is considered that chronic inflammation, caused by the overabundance of nutrients, increases the levels of inflammatory cytokines and immune cells. It has been discovered that adipocytes have an important endocrine role; they synthesize numerous hormones and adipocytokines, such as leptin and adiponectin. High levels of leptons and low levels of adiponectin can activate intracellular signaling pathways involving malignant cells’ development. An important part of cancer development can be attributed to insulin metabolism, insulin-like growth factors, and sex hormones.
Article
Full-text available
There is growing evidence that excess body weight increases the risk of cancer at several sites, including kidney, endometrium, colon, prostate, gallbladder and breast in post‐menopausal women. The proportion of all cancers attributable to overweight has, however, never been systematically estimated. We reviewed the epidemiological literature and quantitatively summarised, by meta‐analysis, the relationship between excess weight and the risk of developing cancer at the 6 sites listed above. Estimates were then combined with sex‐specific estimates of the prevalence of overweight [body mass index (BMI) 25–29 kg/m²] and obesity (BMI ≥30 kg/m²) in each country in the European Union to obtain the proportion of cancers attributable to excess weight. Overall, excess body mass accounts for 5% of all cancers in the European Union, 3% in men and 6% in women, corresponding to 27,000 male and 45,000 female cancer cases yearly. The attributable proportion varied, in men, between 2.1% for Greece and 4.9% for Germany and, in women, between 3.9% for Denmark and 8.8% for Spain. The highest attributable proportions were obtained for cancers of the endometrium (39%), kidney (25% in both sexes) and gallbladder (25% in men and 24% in women). The largest number of attributable cases was for colon cancer (21,500 annual cases), followed by endometrium (14,000 cases) and breast (12,800 cases). Some 36,000 cases could be avoided by halving the prevalence of overweight and obese people in Europe. © 2001 Wiley‐Liss, Inc.
Article
Full-text available
Background The aim of the present study was to evaluate the protective effect of concurrent exercise in the degree of the insulin resistance in mice fed with a high-fat diet, and assess adiponectin receptors (ADIPOR1 and ADIPOR2) and endosomal adaptor protein APPL1 in different tissues. Methods Twenty-four mice were randomized into four groups (n = 6): chow standard diet and sedentary (C); chow standard diet and simultaneous exercise training (C-T); fed on a high-fat diet and sedentary (DIO); and fed on a high-fat diet and simultaneous exercise training (DIO-T). Simultaneously to starting high-fat diet feeding, the mice were submitted to a swimming exercise training protocol (2 x 30 minutes, with 5 minutes of interval/day), five days per week, for twelve weeks (90 days). Animals were then euthanized 48 hours after the last exercise training session, and adipose, liver, and skeletal muscle tissue were extracted for an immunoblotting analysis. Results IR, IRs, and Akt phosphorylation decreased in the DIO group in the three analyzed tissues. In addition, the DIO group exhibited ADIPOR1 (skeletal muscle and adipose tissue), ADIPOR2 (liver), and APPL1 reduced when compared with the C group. However, it was reverted when exercise training was simultaneously performed. In parallel, ADIPOR1 and 2 and APPL1 protein levels significantly increase in exercised mice. Conclusions Our findings demonstrate that exercise training performed concomitantly to a high-fat diet reduces the degree of insulin resistance and improves adipoR1-2/APPL1 protein levels in the hepatic, adipose, and skeletal muscle tissue.
Article
Full-text available
The identification of a complex containing the tumor suppressor LKB1 as the critical upstream kinase required for the activation of AMP-activated protein kinase (AMPK) by metabolic stress was reported in an article in Journal of Biology in 2003. This finding represented the first clear link between AMPK and cancer. Here we briefly discuss how this discovery came about, and describe some of the insights, especially into the role of AMPK in cancer, that have followed from it. In September 2003, our groups published a joint paper [1] in Journal of Biology (now BMC Biology) that identified the long-sought and elusive upstream kinase acting on AMP-activated protein kinase (AMPK) as a complex containing LKB1, a known tumor suppressor. Similar findings were reported at about the same time by David Carling and Marian Carlson [2] and by Reuben Shaw and Lew Cantley [3]; at the time of writing these three papers have received between them a total of over 2,000 citations. These findings provided a direct link between a protein kinase, AMPK, which at the time was mainly associated with regulation of metabolism, and another protein kinase, LKB1, which was known from genetic studies to be a tumor suppressor. While the idea that cancer is in part a metabolic disorder (first suggested by Warburg in the 1920s [4]) is well recognized today [5], this was not the case in 2003, and our paper perhaps contributed towards its renaissance. The aim of this short review is to recall how we made the original finding, and to discuss some of the directions that these findings have taken the field in the ensuing ten years.
Article
Adiponectin (also known as 30-kDa adipocyte complement-related protein; Acrp30) is a hormone secreted by adipocytes that acts as an antidiabetic and anti-atherogenic adipokine. Levels of adiponectin in the blood are decreased under conditions of obesity, insulin resistance and type 2 diabetes. Administration of adiponectin causes glucose-lowering effects and ameliorates insulin resistance in mice. Conversely, adiponectin-deficient mice exhibit insulin resistance and diabetes. This insulin-sensitizing effect of adiponectin seems to be mediated by an increase in fatty-acid oxidation through activation of AMP kinase and PPAR-alpha. Here we report the cloning of complementary DNAs encoding adiponectin receptors 1 and 2 (AdipoR1 and AdipoR2) by expression cloning. AdipoR1 is abundantly expressed in skeletal muscle, whereas AdipoR2 is predominantly expressed in the liver. These two adiponectin receptors are predicted to contain seven transmembrane domains, but to be structurally and functionally distinct from G-protein-coupled receptors. Expression of AdipoR1/R2 or suppression of AdipoR1/R2 expression by small-interfering RNA supports our conclusion that they serve as receptors for globular and full-length adiponectin, and that they mediate increased AMP kinase and PPAR-alpha ligand activities, as well as fatty-acid oxidation and glucose uptake by adiponectin.
Article
Adipose differentiation is accompanied by changes in cellular morphology, a dramatic accumulation of intracellular lipid and activation of a specific program of gene expression. Using an mRNA differential display technique, we have isolated a novel adipose cDNA, termed adipoQ. The adipoQ cDNA encodes a polypeptide of 247 amino acids with a secretory signal sequence at the amino terminus, a collagenous region (Gly-X-Y repeats), and a globular domain. The globular domain of adipoQ shares significant homology with subunits of complement factor C1q, collagen α1(X), and the brain-specific factor cerebellin. The expression of adipoQ is highly specific to adipose tissue in both mouse and rat. Expression of adipoQ is observed exclusively in mature fat cells as the stromal-vascular fraction of fat tissue does not contain adipoQ mRNA. In cultured 3T3-F442A and 3T3-L1 preadipocytes, hormone-induced differentiation dramatically increases the level of expression for adipoQ. Furthermore, the expression of adipoQ mRNA is significantly reduced in the adipose tissues from obese mice and humans. Whereas the biological function of this polypeptide is presently unknown, the tissue-specific expression of a putative secreted protein suggests that this factor may function as a novel signaling molecule for adipose tissue.
Article
AMPK (AMP-activated protein kinase) is a sensor of cellular energy status that appears to have arisen during early eukaryotic evolution. In the unicellular eukaryote Saccharomyces cerevisiae, the AMPK orthologue is activated by glucose starvation and is required for the switch from glycolysis (fermentation) to oxidative metabolism when glucose runs low. In mammals, rapidly proliferating cells (including tumour cells) and immune cells involved in inflammation both tend to utilize rapid glucose uptake and glycolysis (termed the Warburg effect or aerobic glycolysis) rather than oxidative metabolism to satisfy their high demand for ATP. Since mammalian AMPK, similar to its yeast orthologue, tends to promote the more energy-efficient oxidative metabolism at the expense of glycolysis, it might be expected that drugs that activate AMPK would inhibit cell proliferation and and hence cancer, as well as exerting anti-inflammatory effects. Evidence supporting this view is discussed, including our findings that AMPK is activated by the classic anti-inflammatory drug salicylate.