ArticlePDF AvailableLiterature Review

Regulation of insulin secretion and reactive oxygen species production by free fatty acids in pancreatic islets

Authors:

Abstract and Figures

Free fatty acids regulate insulin secretion through metabolic and intracellular signaling mechanisms such as induction of malonyl-CoA/long-chain CoA pathway, production of lipids, GPRs (G protein-coupled receptors) activation and the modulation of calcium currents. Fatty acids (FA) are also important inducers of ROS (reactive oxygen species) production in β-cells. Production of ROS for short periods is associated with an increase in GSIS (glucose-stimulated insulin secretion), but excessive or sustained production of ROS is negatively correlated with the insulin secretory process. Several mechanisms for FA modulation of ROS production by pancreatic β-cells have been proposed, such as the control of mitochondrial complexes and electron transport, induction of uncoupling proteins, NADPH oxidase activation, interaction with the renin-angiotensin system, and modulation of the antioxidant defense system. The major sites of superoxide production within mitochondria derive from complexes I and III. The amphiphilic nature of FA favors their incorporation into mitochondrial membranes, altering the membrane fluidity and facilitating the electron leak. The extra-mitochondrial ROS production induced by FA through the NADPH oxidase complex is also an important source of these species in β-cells.
No caption available
… 
Content may be subject to copyright.
©2011 Landes Bioscience.
Do not distribute.
www.landesbioscience.com Islets 213
Islets 3:5, 213-223; September/October 2011; ©2011 Landes Bioscience
REVIEW REVIEW
Reactive oxygen species (ROS) are important components of
intracellular redox signaling cascades. They are produced by the
metabolism of different substrates such as fatt y acids (FA) and glu-
cose. Mitochondrial production of ROS plays an important role
*Correspondence to: Maria Fernanda Rodrigues Graciano;
Email: mafe@icb.usp.br
Submi tted: 04/15/11; Revised: 06/15/11; Accepted: 06/17/11
DOI: 10.4161/isl.3.5.15935
Free fatty acids regulate insulin secretion through metabolic
and intracellular signaling mechanisms such as induction
of malonyl-CoA/long-chain CoA pathway, production of
lipids, GPRs (G protein-coupled receptors) activation and
the modulation of calcium currents. Fatty acids (FA) are
also important inducers of ROS (reactive oxygen species)
production in β-cells. Production of ROS for short periods
is associated with an increase in GSIS (glucose-stimulated
insulin secretion), but excessive or sustained production of
ROS is negatively correlated with the insulin secretory process.
Several mechanisms for FA modulation of ROS production by
pancreatic β-cells have been proposed, such as the control of
mitochondrial complexes and electron transport, induction of
uncoupling proteins, NADPH oxidase activation, interaction
with the renin-angiotensin system, and modulation of the
antioxidant defense system. The major sites of superoxide
production within mitochondria derive from complexes I and
III. The amphiphilic nature of FA favors their incorporation into
mitochondrial membranes, altering the membrane uidity
and facilitating the electron leak. The extra-mitochondrial
ROS production induced by FA through the NADPH oxidase
complex is also an important source of these species in β-cells.
Regulation of insulin secretion and production
of reactive oxygen species by free fatty acids
in pancreatic islets
Maria Fernanda Rodrigues Graciano,1,* Maíra M.R. Valle,1 Anjan Kowluru,2,3 Rui Curi1 and Angelo R. Carpinelli1
1Departm ent of Physiology and Biop hysics; Institute of Biom edical Sciences; Unive rsity of São Paulo; São Pau lo, SP Brazil;
2Depart ment of Pharmaceutical S ciences; Eugene Appleb aum College of Pharmac y; Wayne State University ; Detroit, MI USA;
3β-Cell Bioch emistry Research L aboratory; John D. D ingell VA Medical Center; De troit, MI USA
Key words: oxidative stress, fatty acids, b-cells, insulin release, NADPH oxidase, hydrogen peroxide, superoxide
Abbreviations: ACS, acyl-CoA synthetase; ARBs, angiotensin II receptor blockers; CAT, catalase; CPT-I, carnitine
palmitoyltransferase I; DAG, diacylglycerol; FA, fatty acids; GPx, glutathione peroxidase; GCLC, glutamylcysteine ligase;
GSIS, glucose-stimulated insulin secretion; GIP, glucose-dependent insulinotropic polypeptide; GLP-1, glucagon-like-peptide
1; GPR, G protein-coupled receptor; HSL, hormone-sensitive lipase; iNOS, induced form of nitric oxide synthase; IP3, inositol
triphosphate; K ATP, ATP sensitive-potassium channels; LC-CoA, long-chain acyl-coenzime A; LDH, lactate dehydrogenase; OAA,
oxaloacetate; LTCC, L-type calcium channels; PIP2, phosphatidylinositol-4,5-bisphosphate; R AS, renin-angiotensin system; ROS,
reactive oxygen species; SNPs, single nucleotide polymorphisms; SOD, superoxide dismutase; TAG, triacylglycerol; TCA cycle,
tricarboxylic acid cycle; UCP, uncoupling protein
in the pancreatic β-cell secretory function.1 However, evidence is
accumulating to suggest that pancreatic islets also produce ROS
through NADPH oxidase activity, and this process is involved
in glucose-stimulated insulin secretion (GSIS).2-4 Production of
ROS for short periods is associated with an increase with GSIS,
but excessive or sustained production of ROS is negatively cor-
related with the insulin secretory process.1,2 Chronic exposure to
relatively high levels of ROS leads to impairment of pancreatic
β-cell function and diabetes.5,6
In this review, the mechanisms involved in fatty acid-medi-
ated insulin secretion and of ROS production in pancreatic islets
and b-cells are discussed. Our main focus is the acute effect of
the FA on islet functions and on the signaling pathways involved
in the amplification of the glucose-induced insulin secretion.
These early signals are also important to elucidate the chronic
effects of FA in conditions such as insulin resistance, obesity and
type 2 diabetes.
Mechanisms by which FA Stimulate Insulin Secretion
FA play an important role in β-cell function. FA deprivation
in islet reduces GSIS,7,8 which is restored by exogenous free
FA.9 Short time exposure (1 h) of pancreatic islets to FA aug-
ments GSIS;10 -13 however, if chronically maintained in excess,
saturated FA reduce insulin biosynthesis14 and secretion.15-17
The potency of FA to promote glucose-induced insulin release
increases with the chain length and decreases with the degree
of unsaturation.18,19 There are pronounced differences in cell
effects of saturated and unsaturated FA, with the latter display-
ing a tendency to promote cell viability under conditions which
otherwise would be cytotoxic. The mechanisms involved in this
cytoprotective action remain under intense investigation, but
©2011 Landes Bioscience.
Do not distribute.
214 Islets Volume 3 Issue 5
synthetase (ACS) and enter the mitochondria, being oxidized via
the β-oxidation pathway for ATP production23 (Fig. 1).
Pyruvate is converted into acetyl-CoA via the pyruvate dehy-
drogenase complex (PDH) or to oxaloacetate (OAA) via pyruvate
carboxylase (PC), in general with approximately equal amounts
entering each metabolic flux in the mitochondria.24, 25 The cyto-
solic substrate flux via lactate dehydrogenase (LDH) is limited by
the low activity of this enzyme in pancreatic islets.26 The cataple-
rotic activity of PDH, in association with the flux through the
anaplerotic PC, links glycolysis with OAA/citrate synthesis and,
in the fed state or in high glucose concentrations, with the synthe-
sis of malonyl-CoA. Factors that lead to PDH inhibition, there-
fore, favor acetyl-CoA production via β-oxidation of long-chain
FA. Therefore, intracellular FA homeostasis, in part mediated via
the malonyl-CoA nutrient-sensing mechanism, plays a key role
the rapid inhibition of caspase 3 and the reduction of endoplas-
mic reticulum stress by the unsaturated FA could be involved.20
Malonyl-CoA/LC-CoA (long-chain acyl-Coenzyme A)
metabolic pathway. Circulating levels of FA control GSIS as
occurs in fasted rats. Under this condition, b-cells are not able
to convert glucose into malonyl-CoA, possibly as a consequence
of low pyruvate dehydrogenase activity.21 Malonyl-CoA inhibits
mitochondrial long-chain fatty acid uptake by inhibiting carni-
tine palmitoyltransferase I (CPT-I), thereby promoting FA re-
esterification. As a result, there is no suppression of CPT-I activ-
ity in fasted rats (which avoids the rise of the cytosolic acyl-CoA
concentration) and the insulin secretory process is suppressed.
This condition renders the b-cell dependent upon a high exter-
nal FA supply to counterbalance the deficit.22 When the glu-
cose level is low, FA are converted into LC-CoAs by acyl-CoA
Figure 1. Mechanisms by which FA stimulate insulin secretion. Pyruvate can be converted to acetyl-CoA via the pyruvate dehydrogenase complex
(PDH) or to oxaloacetate (OAA) via pyruvate carbox ylase (PC) in pancreatic islets. The cataplerotic activity of PDH, in association with ux through the
anaplerotic enzyme PC, links glycolysis with OAA/citrate synthesis and, in the fed state or in high glucose concentrations, with the synthesis of
malonyl-CoA. Malonyl- CoA inhibits mitochondrial long-chain fatty acid uptake at the carnitine palmitoyltransferase I (CPT-I) level, and thereby
promotes FA re-esterication. High glucose concentrations raise malonyl- CoA levels and decrease FA oxidation, which leads to LC-CoA accumulation.
LC-CoA may additionally be esteried to diacylglycerol (DAG) and triacylglycerol (TAG) in the presence of glycerol-3-phosphate provided by glucose
metabolism. DAG formation and PKC activation may synergize with the classic pathways of insulin secretion—KATP closure and calcium inux—to
promote full insulin secretion. GPR40 activation by FA stimulates the Gαq-PLC (phospholipase C) signaling pathway, leading to calcium release from
endoplasmic reticulum stores and to DAG production. Endogenous lipolysis by the action of hormone-sensitive lipase (HSL) regulates insulin secretion
through the generation of FA and other lipid-signaling molecules. ACS, acyl-CoA synthetase; LDH, lactate dehydrogenase; LTCC, L-type calcium chan-
nels; PL, phospholipids.
©2011 Landes Bioscience.
Do not distribute.
www.landesbioscience.com Islets 215
(arginine to histidine) at the position 211 in the third intracellu-
lar loop of the receptor that was reported to cause enhanced insu-
lin secretion in Japanese men.40 The other SNP also produces an
amino acid substitution (Gly180Ser),41 and the less frequent vari-
ant (Ser180) has been associated with elevated body-mass index
in Europeans and a reduction of insulin secretion following an
oral glucose load. Similarly, insulin secretion was also reduced in
this group during an oral lipid load. Studies in transfected HeLa
cells with the Ser180 variant showed low cytosolic calcium levels
compared to those expressing the Gly180 receptor in response to
FA load .
Isolated rat pancreatic islets export a substantial amount of
FA to the incubation medium either in the absence or presence of
glucose. After incubation for 1 h, the addition of 5.6 mM glucose
raises the medium content of palmitic and stearic acids compared
with islets incubated in the absence of glucose. Martins et al con-
cluded that the synthesis and release of saturated and, to a lesser
extent, unsaturated FA from glucose-exposed islets represent an
amplification pathway for insulin secretion.42 The FA released
through triglyceride/FFA cycling and partially secreted from
pancreatic islets may activate the GPR40 pathway via autocrine
and/or paracrine mechanisms43 (Fig. 1)
The FA are supplied to pancreatic b-cells from intracellular
triglycerides and phospholipids and from plasma FFA and lipo-
proteins.43 The FFA originated from plasma triglycerides can
access β-cells by the action of lipoprotein lipases.44 In addition,
the action of hormone-sensitive lipase (HSL) in b-cells reinforces
the concept that endogenous lipolysis participates in the regula-
tion of insulin secretion through generation of FA or other lipid-
signaling molecules.45 In fact, the islet triglyceride stores may play
an important role in GSIS via HSL-lipolysis (Fig. 1). The HSL
KO mice present reduced GSIS both in vivo and in isolated islets.
Increases in glucose concentration induce HSL expression, which
occurs concomitantly with an augmentation of basal insulin
secretion.46,47 These observations indicate that the FA exported
from the β-cells may act as signaling molecules in GSIS.
Acylation of specific proteins leads to the amplification of
insulin secretion. Myristoylation and palmitoylation are involved
in directing proteins to appropriate membrane sites, thus stabiliz-
ing protein-protein interactions and regulating certain enzyme
activities in the mitochondria.48 Various proteins can be acylated:
α subunits of G-proteins, tyrosine kinases receptors and proteins
involved in the regulation of ion channel activity and exocy-
tosis.29,4 9,50 FA have also been proposed to regulate ionic chan-
nels in pancreatic b-cells.29 ,51 Cerulenin, an inhibitor of protein
acylation, inhibits the augmentation of calcium-induced insulin
release promoted by palmitate.49
Calcium currents. In Chinese hamster ovary cells express-
ing human, mouse and rat GPR40, saturated FA with lengths
ranging from C12 to C16 and C18 and C22 unsaturated FA all
present calcium influx-inducing activities.31 FA increase calcium
influx in β-cells by a mechanism dependent on PLC activa-
tion;52,53 and L-type calcium channels (LTCC).52-54 The palmitic
acid activation of GPR40 is dependent upon high glucose con-
centrations. This suggests that GPR40 signaling requires cal-
cium influx through LTCC that are pre-activated by glucose,
in pancreatic β-cell function (Fig. 1). These processes are early
events in the induction of insulin secretion in pancreatic β-cells.27
High glucose concentrations raise malonyl-CoA levels and
decrease FA oxidation, which leads to LC-CoA accumulation.
LC-CoA facilitates the fusion of secretory granules with the
β-cell plasma membrane, thus promoting insulin secretion.28
The nonmetabolizable analogue 2-bromopalmitate (an inhibitor
of CPT-I, which inhibits fatty acid oxidation and is expected to
increase cytosolic long chain acyl-CoAs) also stimulates insulin
release.29 LC-CoA may additionally be esterified to diacylglyc-
erol (DAG) and triacylglycerol (TAG) in the presence of glycerol-
3-phosphate provided by glucose metabolism. DAG formation
and PKC activation may synergize with the classic pathways of
insulin secretion—ATP sensitive-potassium channels (KATP )
closure and calcium influx—to promote full insulin secretion30
(Fig. 1).
GPR40 activation and production of lipid-signaling mol-
ecules. The GPR (G protein coupled receptor) isoforms 40, 41
and 43 are highly homologous but differ in substrate specific-
ity and tissue distribution. They were formerly known as orphan
receptors, however, it was found that FA are agonists of these
receptors.31 GPR41 and GPR43 are specifically activated by
the short-chain FA acetic, propionic and butyric acids. mRNA
expression of both receptors was detected in mouse pancreatic
islets as well as in insulin-producing MIN6 cells.32
Long-chain FAs can signal directly via the FA receptor GPR40.
This Gαq membrane receptor is highly expressed in pancreatic
b-cells and potentially activated by the most prevalent FA in
plasma.31,33,34 Taste buds35 and intestinal K and L cells express this
receptor,3 6,37 where they are able to signal fat ingestion, contributing
for the secretion of the incretins glucagon-like-peptide (GLP-1) and
glucose-dependent insulinotropic polypeptide (GIP). Therefore,
FA can directly stimulate insulin secretion through GPR40 in
β-cell surface and, indirectly, potentiate incretin secretion.
FA binding to GPR40 activates a heterotrimeric G protein,
containing the α subunit of the Gq protein, which is known to
stimulate PLC activity. This enzyme converts phosphatidylinosi-
tol-4,5-bisphosphate (PIP2) into DAG and inositol triphosphate
(IP3). DAG is a well-known PKC activator and IP3 participates
in calcium extrusion from the endoplasmic reticulum (Fig. 1).
The RNAi-mediated knockdown of GPR40 expression in
MIN6 cells abolishes the increase of insulin secretion induced
by linoleic and γ-linolenic acids.31 The relevant participation of
GPR40 for β-cell function has also been shown in vivo. Latour
and collaborators (2007) showed that insulin secretion response
to intralipids is reduced by approximately 50% in GPR40 knock-
out (KO) mice.38 Lan and collaborators (2008) also demon-
strated a reduction of insulin secretion by 50% in response to
elevated seric fatty acid concentrations in GPR40 KO mice.39
These results show that the signaling induced by GPR40 pathway
and the intracellular fatty acid metabolism are complementary
for the fatty acid induction of insulin secretion.
At least two single nucleotide polymorphisms (SNPs) with
functional effects, and located in the coding region of the
human GPR40 gene, have been described in references 40 and
41. One of these SNPs leads to an amino acid substitution
©2011 Landes Bioscience.
Do not distribute.
216 Islets Volume 3 Issue 5
electrons for entry at complex I, NADH dehydrogenase, which
is also known as NADH ubiquinone oxidoreductase; or complex
II, succinate dehydrogenase, through acetyl-CoA metabolism in
the TCA cycle. In addition, FA metabolism provides electrons
through the electron-transport flavoprotein via FADH2, which is
a product of β-oxidation, independent of the TCA cycle.72
The major sites of superoxide production within mitochon-
dria derive from complex I and III.73 Complex I superoxide arises
from bound flavin reduced FMNH2 by NADH and is released
nearly exclusively to the matrix side of the inner membrane.
Complex III is the cytochrome bc1 complex and the superoxide is
generated during Q cycle, wherein coenzyme Q undergoes redox
cycling through a reactive semiquinone species72 and generates
superoxide to both the matrix and outward to the intermembrane
and extramitochondrial space74 (Figs. 2 and 3).
Superoxide production has been demonstrated to participate
as a signal to acutely enhance insulin secretion by various stimuli,
such as glucose and FA.1,3,4,75 More details of this signaling pro-
cess induced by FA will be discussed in the section dealing with
NADPH oxidase.
Modulation of the electron transport. A general condition
that favors mitochondrial superoxide generation is the highly
reduced state of the electron carriers at specific sites. This enables
the leak of electrons out of the enzymatic electron transport
route (respiratory chain). Therefore, mitochondrial ROS pro-
duction depends on the redox state of electron-donating centers
of complexes I and III, such as flavin mononucleotide (FMN),
Fe-S clusters and Q binding sites76 (Fig. 3). Enhanced cellular
concentrations of LC-CoA, as observed in obesity, may increase
transmembrane potential and shift the redox state of coenzyme
Q toward more reduced values,77 thus promoting mitochondrial
superoxide production.
FA interact with components of the respiratory chain, thereby
inhibiting electron transport.78 β-oxidation of FA supports
reverse electron transport.79 In this mechanism, electrons are
transferred from FA to the flavoprotein (electron transfer flavo-
protein) and enter the respiratory chain at the coenzyme Q level,
a reaction mediated by flavoprotein-quinone oxidoreductase, thus
increasing the superoxide production associated to the complex I
(Fig. 3).
FA can interact with the respiratory chain by binding to
cytochrome c in complex III,80 which interrupts the electron
transport and contributes to enhance ROS production. In this
way, FA might induce ROS production due to the depletion
of cytochrome c from mitochondria, thereby interrupting the
electron flow from complex III to complex IV (Fig. 3). The
amphiphilic nature of FA favors their incorporation into phospho-
lipid bilayers or mitochondrial membranes and alters the mem-
brane fluidity. This may facilitate the electron leak from the inner
mitochondrial membrane and one-electron reduction of O2.79
Ceramides are amides of long-chain FA and sphingosine,
essential building blocks of sphingolipids and components of bio-
logical membranes. Ceramides have been recognized as impor-
tant signaling molecules in cell proliferation, differentiation
and, in particular, cell death. They stimulate respiratory chain-
associated ROS production by depletion of mitochondrial
resulting in increased calcium influx and augmented insulin
secretion.53
The outward voltage-gated potassium currents are required
for the β-cells membrane repolarization after glucose stimulation,
limiting calcium influx and insulin secretion.55 Linoleic acid acti-
vation of GPR40 reduces the activity of the voltage-gated rectifier
potassium channels, leading to reduced potassium conductance
and prolonging the opening of the LTCC. This, in fact, prolongs
the membrane depolarization period during action potential fir-
ing, elevating calcium influx and amplifying the insulin secretion
induced by FA.56
Chronic effects of FA. Several mechanisms are involved
in FA-induced impairment of insulin secretion. One of these
mechanisms involve excessive ROS generation by mitochon-
drial and extramitochondrial sources such as NADPH oxidase,
accompanied by reduction of antioxidant defense via glutathi-
one depletion;57,58 FA-mediated NFκB activation;59 induction of
iNOS (the induced form of nitric oxide synthase, encoded by the
NOS-2 gene in humans), a process also associated with GPR40
signaling;60 FA-derived ceramide production involved in β-cell
death; 61 endoplasmic reticulum stress;62 and induction of mito-
chondrial dysfunction.63 The metabolic effects of FA are also
involved in b-cell dysfunction and induce the reduced glucose
oxidation. In this process, the increased NADH production via
FA β-oxidation inhibits islet pyruvate dehydrogenase activity,
leading to a decrease in the conversion of pyruvate into acetyl
CoA and promoting a reduction in glucose oxidation.21 As men-
tioned above, we focused here on the short-time effects of FA.
Detailed information on chronic effects can be obtained in other
sources.21,57,60,62-64
Mitochondrial ROS Production Induced by FA
Biologically, ROS include the superoxide radical O2
•-, hydrogen
peroxide H2O2, and the hydroxyl radical, OH*. At the physio-
logical pH, superoxide spontaneously dismutates, which is more
efficiently performed by superoxide dismutase (Mn-SOD in
the mitochondrial matrix, Cu/Zn-SOD in the cytosol) to form
H2O2. H2O2 is converted to oxygen and water by catalase (CAT)
and by glutathione peroxidase (GPx).65
ROS are involved in pathological conditions such as diabe-
tes,6 cardiovascular diseases66-68 and neurodegeneration familial
amyotrophic lateral sclerosis.69 ROS also play a role in physiologi-
cal processes such as vascular smooth muscle function,70 insulin
signaling pathway71 and insulin secretion,1-3 by acting as second
messengers. Mitochondria are generally considered sources of
cellular ROS, but are also recognized as organelles with a high
capacity of antioxidative defense through Mn-superoxide dis-
mutase (Mn-SOD), matrix glutathione and glutathione peroxi-
dase (Fig. 2).
Substrates for the tricarboxylic acid cycle (TCA cycle) enter
the mitochondrial matrix through pyruvate dehydrogenase,
carrier proteins or one of multiple shuttle mechanisms. The
metabolism of the substrates results in electron donation to spe-
cific complexes or sites. Fatty acyl-CoAs enter through CPT-I
for b-oxidation in pancreatic β-cells. FA oxidation generates
©2011 Landes Bioscience.
Do not distribute.
www.landesbioscience.com Islets 217
the electron acceptor that leads to formation of hydrogen perox-
ide. This reaction is catalyzed by specific peroxisomal acyl-CoA
oxidase isoforms in rat and humans. So, peroxisomal FA metabo-
lism may contribute to the production of hydrogen peroxide also
in pancreatic islets.83
Uncoupling proteins. The mitochondrial membrane poten-
tial is generated by charge differences across the inner membrane,
and that electrical potential is required for ATP production. This
potential depends on substrate utilization and is generated by
proton pumping at complexes I, III and IV and offset by proton
transfer in the opposite direction, a process called proton leak.
A prominent amount of proton leak is mediated by the uncou-
pling proteins (UCPs).84 UCP2 is the most ubiquitous form
cytochrome c and by direct inhibition of electron transport.81
Other mechanisms are associated to the mitochondrial ROS
production by FA, such as the interaction with the antioxidant
enzymes, as will be later discussed in the section on antioxidant
defenses.
β-oxidation of FA occurs in mitochondria and peroxisomes
in higher eukaryotes. Short and medium chain (C4–C8) FA are
exclusively β-oxidized in the mitochondria, whereas C10–C16
FA are β-oxidized in mitochondria and peroxisomes (C14–C16).
Long and very long-chain (C17-C24) FA are handled preferen-
tially by peroxisomes.82 In mitochondria, the electrons are trans-
ferred to flavin adenine nucleotide (FAD) and nicotinamide
adenine dinucleotide (NAD+); however, in peroxisomes, O2 is
Figure 2. ROS generating systems and antioxidants defenses in pancreatic islets. (1) Fatty acid modulation of NADPH oxidase catalyzes the superoxide
production in phagocy tic cells and pancreatic islets. DAG and calcium increases might be involved in PKC activation, which phosphorylates NADPH
oxidase subunits, inducing the translocation of the cytoplasmic subunits to the plasma membrane core for the NADPH oxidase holoenzyme assembly.
(2) The extracellular SOD (EC-SOD) dismutates superoxide to H2O2 that can diuse through aquaporin channels (AQP) in the plasma membrane to elicit
an intracellular signaling response. Superoxide can also initiate intracellular signaling by permeation of the plasma membrane through anion channels
(Cl-channel-3, ClC-3). (3) Antioxidant defenses involve the cytosolic Cu/Zn-SOD (superoxide dismutase), the mitochondrial Mn-SOD and glutathione
peroxidase (GPx) and catalase (CAT) in cy tosolic, mitochondrial and peroxisomal compar tments. (The size of the letters expresses dierential expres-
sion of the antioxidant defenses in b-cells). Also, the GSH/GSSG (glutathione reduced/glutathione oxidized) ratio is an important antioxidant defense
system in b-cells. Short and medium chain FA are exclusively b-oxidized in the mitochondria (4), whereas long-chain FA are b-oxidized in the mito-
chondria and the peroxisomes (5) and very long-chain FA are handled preferentially by peroxisomes. In mitochondria, the electrons are transferred to
FAD; however, in peroxisomes, O2 is the electron acceptor that leads to formation of hydrogen peroxide. The complexes I and III are the main source of
mitochondrial superoxide, and UCP2 acts as a negative regulator of mitochondria-derived ROS production.
©2011 Landes Bioscience.
Do not distribute.
218 Islets Volume 3 Issue 5
A signaling role of UCP2 could be important during the
period of low glucose concentration, such as during sleep, when
FA oxidation is enhanced, resulting in superoxide production.94,95
As a consequence, the mild uncoupling by UCP2 guarantees
that the β-cell responds properly to the lack of glucose, as it will
reduce insulin secretion despite the availability of an adequate
substrate for ATP production.96 Moreover, UCP2 KO protects
mice from fatty acid-induced impairment in GSIS.97 In physi-
ological conditions with higher glucose and fatty acid concentra-
tions, as after a meal, the superoxide induction of UCP decreases
the proton motive force, which reduces potential oxidative dam-
age but impairs insulin secretion.96
The transcription of the UCP2 gene is highly inducible under
conditions of oxidative stress, such as those induced by lipopoly-
saccharide, free FA and high-fat diet.98-10 0 Superoxide and the lipid
peroxidation product 4-hydroxy-2-nonenal both activate UCP2,
increasing proton conductance in the mitochondrial inner mem-
brane.79 A high fat ketogenic diet increases UCP2 mRNA and pro-
tein levels and reduces ROS production in the brain.101 Conversely,
a low fat diet given to immature rats reduces UCP2 levels and
increases ROS production and seizure-induced excitotoxicity.102
Thus, the ketogenic diet may be neuroprotective by diminishing
ROS production through activation of UCP2 in the brain.101
Modulation of NADPH Oxidase Activity by FA
The NADPH oxidase complex is well characterized in phago-
cytic cells, where it consists of the cytosolic components p47PHOX,
p67PHOX, p40PHOX, a low molecular weight G-protein, Rac 1 or
Rac 2, and the membrane-associated cytochrome b558, includ-
ing g p91PHOX or NOX2 (the catalytic subunit, with the trans-
membrane redox chain) and p22PHOX. In phagocytic cells, PKC
activates NADPH oxidase by a phosphorylation-dependent acti-
vation of p47PHOX, p67PHOX, and/or Rac.103 In these cells, activa-
tion of the NADPH oxidase complex requires translocation of the
cytosolic components to the plasma membrane and their associa-
tion with gp91PHOX/p22PHOX. p47PHOX phosphorylation leads to a
conformational change in this molecule that allows the interaction
with p22PHOX. The p47PHOX subunit translocation to the plasma
membrane leads the p67PHOX into contact with gp91PHOX, bring-
ing the p40P HOX subunit to the complex. Finally, the GTPase Rac
interacts with gp91PHOX. This assembled complex is active and pro-
duces superoxide by electron transfer from NADPH to oxygen.103
Expression of the components of the NADPH oxidase com-
plex and its homologues (NOX1, NOX4, NOXO1 and NOXA1)
has been shown in pancreatic islets.104,105 We demonstrated the
induction of p47PHOX translocation to plasma membrane by a
glucose stimulus in pancreatic islets.104 Furthermore, NADPH
oxidase has been identified in caveolae and lipid rafts in various
cell types, such as coronary endothelial cells and vascular smooth
muscle cells.106-108 These structures are cholesterol and sphingo-
lipid-rich plasma membrane microdomains where multiple sig-
naling molecules, including GPRs, tyrosine kinase receptors,
PKC and G proteins are localized, promoting the compartmen-
talization of signaling.109
and it is also expressed in pancreatic β-cells.85 UCP1 dissipates
caloric energy as heat, by uncoupling mitochondrial respiration
from ATP production. However, UCP2 is not a physiologically
relevant “uncoupling protein” like UCP1 and does not contribute
to adaptive thermogenesis.86
UCPs are also able to promote the export of FA from mito-
chondria to the cytosolic compartment, thus protecting against
FA overload. For example, the antioxidative activity of UCP3
decreases the matrix content of FA.87 UCPs transfer fatty acid
peroxides from the inner to the outer leaflet of the inner mito-
chondrial membrane, thus extruding these highly toxic peroxida-
tion products.88
UCPs can be activated by extramitochondrially generated
superoxide89 and by superoxide generated in the mitochondrial
matrix under nonphysiological90 and physiological conditions.91
UCP2 in β-cells decreases GSIS and the removal of UCP2 results
in higher ATP levels and improved insulin secretion in mice
islets.85 UCP2 functions as a negative regulator of mitochondria-
derived ROS production. UCPs may respond to overproduction
of matrix superoxide by catalyzing mild uncoupling, which low-
ers proton motive force and decreases superoxide production from
electron transport chain, attenuating superoxide-mediated dam-
age at the cost of slightly lowered efficiency of oxidative phos-
phorylation.92 UCP2 KO mice, on three highly congenic strain
backgrounds, exhibit increased oxidative stress and decreased
GSIS.93 Therefore, chronic absence of UCP2 may disrupt the
adaptive response to oxidative stress.
Figure 3. ROS generation in the mitochondrial elec tron transport chain
and the modulation by FA. The major sites of superoxide production
within mitochondria derive from complex I and III of the respiratory
chain. FA inhibit the elec tron transport within complexes I and III, thus
facilitating the electron leak and enabling one-electron reduction of
oxygen to superoxide. FA might induce ROS produc tion due to the
depletion of c ytochrome c from mitochondria, thereby interrupting the
electron ow from complex III to complex IV, increasing the reduction
state of upstream electron carriers. Fum, fumarate; Succ, succinate;
R FeS, Rieske iron-sulfur protein; cy t c1, cyt c, cyt a/a3: the respective
cytochromes. Interactions of FA in electron transport are indicated by
dotted lines. Solid lines show the direction of electron transfer. Adapted
from P. Schönfeld, L. Wojtczak (2008).79
©2011 Landes Bioscience.
Do not distribute.
www.landesbioscience.com Islets 219
insulin secretion in high glucose concentrations.2,3 In cardiac
myocytes, endotelin-1 increases ROS production via NADPH
oxidase, which plays a key role in calcium influx through L-type
calcium channels. This probably occurs due to the redox modi-
fication of cysteine residues on the cardiac L-type calcium
channel.112
Chronic effects of palmitic and oleic acids inducing ROS
production are restored by suppression of gp91PHOX,113 and the
chronic effect of palmitate-induced reduction of GSIS is pre-
vented by incubation of mouse islets with the NADPH oxidase
inhibitor apocynin.114 On the other hand, the concomitant incu-
bation of the unsaturated FA arachidonic acid with palmitic acid
dose-dependently reduces the saturated FA induction of ROS and
NO production in BRIN-BD11 cells. Arachidonic acid reduced
the expression of iNOS, NFκB and p47PHOX and improved GSH/
GSSG ratio in palmitic acid-treated cells.115
As mentioned, fatty acid-induced ROS production for short
time stimulates insulin secretion and, in chronic stimulation,
reduces GSIS. There is possibly an optimal range of ROS content
to maintain β-cells function, and a variation of the redox state
due to excessive ROS formation or marked ROS reduction results
in reduced insulin secretion.
The renin-angiotensin system and NADPH oxidase. The
pancreatic islets are exposed not only to systemic but also to
locally produced components of the renin-angiotensin system
(RAS). R AS, in conditions such as obesity and diabetes mel-
litus type 2, is inappropriately upregulated. Angiotensinogen,
the angiotensin converting enzyme (ACE), and the angiotensin
receptors 1 and 2 (AT-1 and AT-2) have all been found in rodent
pancreatic islets.116-118 In the human pancreas, renin precursors
and AT-1 have been found in β-cells, as well as in endothelial
cells of the pancreatic vasculature.119
Increasing concentrations of angiotensin II impairs GSIS in
a dose-dependent manner in mouse islets and this effect is abol-
ished by losartan, an AT-1 receptor antagonist.116 ,12 0 Angiotensin
II induces a dose-dependent superoxide generation via NADPH
oxidase activation and increases protein and mRNA lev-
els of NADPH oxidase subunits (p47PH OX and g p91PHOX).121
Hyperglycemia per se can activate RAS in human islets, and
under high glucose concentrations, angiotensin converting
enzyme inhibitors exert beneficial effects on β-cell, improv-
ing insulin secretion and reducing nitrotyrosine and p22PHOX
mRNA levels.122 Thus, RAS triggers the production of ROS via
NADPH oxidase complex in pancreatic islets, regulating insulin
secretion.
Obesity-induced diabetes mellitus type 2 in db/db mice has
been related to b-cell dysfunction, likely via activation of pan-
creatic RAS and upregulation of AT-1 receptors in the pancreas.
Angiotensin II receptor blockers (ARBs) increase both insulin
production and secretion. Furthermore, hyperglycemia, glucose
intolerance, and the onset of diabetes are delayed by ARBs, with-
out affecting insulin resistance.123
ARBs have received a great deal of attention as therapeu-
tic tools for obesity-related metabolic disorders. The increased
expression of p22PHOX a nd gp91P HOX is correlated with increased
oxidative stress in islets of the type 2 diabetes models, OLEFT
In phagocytes, gp91PHOX can also be activated within the
granules without the need of fusion with surface membranes.
However, all NOX family members are transmembrane proteins
that transport electrons across biological membranes to reduce
oxygen to superoxide. NOX2 is a transmembrane redox chain
that connects the electron donor, NADPH, on the cytosolic
side of the membrane with the electron acceptor, oxygen, on
the outer side of the membrane. It transfers electrons through
a series of steps involving a FAD and two assymetrical hemes
found in transmembrane domains III and V. Electrons are ini-
tially transferred from NADPH to FAD, a process that is regu-
lated by the activation domain of p67PHOX . A single electron is
transferred from the reduced flavin FADH2 to the iron center of
the inner heme. The inner heme must donate its electron to the
outer heme before the second electron can be accepted from the
now partially reduced flavin, FADH. Oxygen must be bound to
the outer heme to accept the electron.103 Overall, g p91PHOX trans-
fers electrons from intracellular NADPH to extracellular oxygen,
generating superoxide anion. The latter is dismutated to H2O2
by the extracellular SOD (EC-SOD). H2O2 can diffuse through
aquaporin channels in the plasma membrane to elicit intracellu-
lar signaling responses. Superoxide can also initiate intracellular
signaling by penetrating the cell membrane through anion chan-
nels (Cl-channel-3, ClC-3) 110 (Fig. 2).
Short-time exposure to palmitate in the presence of low glu-
cose concentration and pro-inflammatory cytokines increases
superoxide production through NADPH oxidase activation in
insulin-secreting cells and pancreatic islets.75 Acute palmitate
induction of superoxide production in islets was demonstrated
to be dependent on the activation of PKC and NADPH oxi-
dase, causing p47PHOX translocation to plasma membrane and
upregulation of the p47PHOX protein content and of the p22PHOX,
gp91PHOX and p47P HOX mRNA levels. Palmitic acid oxidation
contributes to its induction of superoxide production in the
presence of 5.6 mM glucose, as observed in experiments per-
formed with a CPT-I irreversible inhibitor, etomoxir.4 This is
probably a consequence of the electron transfer from FA to the
flavoprotein that enters the respiratory chain at the coenzyme Q
level and, through the reverse electron transport, increases the
superoxide production associated to the complex I, as previously
discussed.
On the other hand, in pancreatic islets of Wistar rats fed with
a high fat diet during 3 months, glucose metabolism is increased,
insulin secretion elevated in the presence of high glucose lev-
els, protein expression of NADPH oxidase subunits is reduced
and both ROS production and apoptosis are diminished.111 The
downregulation of the NADPH oxidase complex has a role in
the compensation of the deleterious effect of lard, promoting the
equilibrium of the cellular redox state (as demonstrated by the
absence of differences in oxidative stress markers)111 to maintain
insulin secretion and to avoid the initial event of insulin resis-
tance in obese rats.
FA stimulation of insulin secretion in the presence of high
glucose concentration is reduced by inhibition of NADPH oxi-
dase activity.4 In fact, NADPH oxidase inhibition has been
associated with reduced glucose oxidation, calcium influx and
©2011 Landes Bioscience.
Do not distribute.
220 Islets Volume 3 Issue 5
not only for rodent pancreatic islets but also for the rat β-cell
li nea ge RINm5F.135 In 2004, the level of expression of glutamyl-
cysteine ligase (GCLC), the enzyme that regulates the de novo
synthesis of glutathione, was shown in pancreatic islets to be
similar to other tissues.136
The reasons for this islet low antioxidant defense are still
being debated. Why such important cells are so vulnerable to
reactive oxygen species? A co-evolution of β-cells, brain cells and
corticosteroids and cortisol receptors may be at the origin of this
peculiarity of the endocrine pancreas.137 The initial clues were
obtained by the fact that both human and rodent females pres-
ent low antioxidant capacity in pancreatic islets.138,139 The fact
that during pregnancy females present insulin resistance provides
an indication as to why evolution maintained low levels of anti-
oxidant enzymes in β-cells.137 The insulin resistance progresses
slowly during pregnancy, supplying extra circulating glucose
according to the fetus needs.140 The molecular mechanism that
prevents the excess of insulin secretion in response to periph-
eral insulin resistance is augmented ROS content.137 The brain
needs a constant supply of glucose and cannot store it.141 In the
same line of reasoning, cortisol and corticosteroid induce insulin
resistance to guarantee the fuel supply necessary to the central
nervous system during stress.137 However, ROS are involved in
insulin secretion, especially H2O2 at low concentrations.1,3,142 In
this regard, there must be other physiological reasons for the low
antioxidant capacity of pancreatic β-cells.
Besides the fact that insulin-producing cells have poor anti-
oxidant defenses and that this fact must have an adaptive mean-
ing, the environmental conditions offered by the current style
of life causes chronic stress, glucolipotoxicity, oxidative stress
and consequent β-cell failure. Chronic high glucose levels
(30 mM) augment the expression of MnSOD in β-cells.135
Chronic treatment with the most common free FA in our
diet, palmitate and oleate, modulates the levels of antioxidant
enzymes in pancreatic islets. Vandewalle and coworkers found
that the incubation of human pancreatic islets with palmitate
0.33 mM for 48 h diminishes the mRNA levels of GPx and
SOD. The treatment with 1 μM of rosiglitazone abolished this
effect; however, it was not dependent on PPAR-γ.143 In human
pancreatic islets, Bikopoulos and co-workers found that chronic
incubation (48 h) with 0.4 mM oleate augments mRNA levels
of catalase, metallothionein 1F and sequestosome 1, molecules
involved in the redox balance.144 The differences between treat-
ments may be due to the different actions of the two free FA.
In opposition to palmitate, which presents deleterious effects,
oleate is considered less damaging and even protective for islet
cells.5
Concluding Remarks
Several mechanisms for FA induction of ROS production by
pancreatic β-cells (mitochondrial complexes and electron trans-
port, NADPH oxidase, antioxidant defenses) were discussed
herein. The balance between ROS production and consump-
tion defines an optimal range of ROS concentration to maintain
b-cell function. The high susceptibility of pancreatic islets to
rats and db /db mice. The inhibitory effect of AT1 ARB val-
sartan on the expression of these components in islets of
db/db mice occurs concomitantly with a reduction of oxidative
stress a nd preservation of insul in content.124 Telmi sarta n, another
ARB, improved insulin sensitivity and reduced the incidence of
type 2 diabetes in patients with hypertension.125 In rats fed a
high-fat and high-carbohydrate diet, telmisartan was shown to
reduce weight gain and significantly reduce the levels of plasma
glucose, insulin and triglycerides.126 The ARB irbesartan
attenuates oxidative stress in islets of Zucker diabetic fat rats
and this process may be a consequence of RAS blockade,
inhibiting NADPH oxidase activation.127 The ARBs candes-
artan and telmisartan both decrease palmitate-induced ROS
accumulation in MIN6 cells and in mouse islets, an effect
dependent on PKC and NADPH oxidase activation.128,129 Also,
the treatment with candesartan recovered palmitate-induced
decrease of intracellular insulin content.129 FA and RAS activa-
tion modulate common intracellular signals, such as NADPH
oxidase activation, that control b-cell function. This could
explain the positive effects of ARBs in high-fat diet and dia-
betic conditions.
Free FA modulation of Rac1 activation by Tiam1. The
activation of small Rho-GTPase proteins, including Rac1,
depends on the change of the bound GDP to GTP. This pro-
cess is regulated by several factors, such as Rho GDIs (GDP
dissociation inhibitor), GEFs (guanine nucleotide exchange
factors) and GAPs (GTPase-activating proteins).69 Several
proteins play these three roles in different situations and tis-
sues. In pancreatic islets, the main GEF responsible to activate
rac1 is Tiam1.130 Tiam1 (T-lymphoma invasion and metastasis)
is a GEF involved in many cellular process, but in β-cell it is
associated with GSIS and activation of NADPH oxidase com-
plex.130,131 As NOX2 is a ROS producer, the activity of Tiam1,
and consequently of Rac1, can be related to β-cells dysfunc-
tion caused by oxidative stress. The incubation of the β-cell line
INS-1 832/13 with palmitate causes activation of Tiam1/Rac1
and ROS production by NOX2. This palmitate action is medi-
ated by ceramide. The use of a pharmacological inhibitor of
Tiam1 (NSC23766) reduces this response, and the dysfunction
generated by palmitate in pancreatic b-cells possibly occurs via
a Tiam1-Rac1 signaling pathway.132
Modulation of the Antioxidant Enzyme Activities
in Pancreatic Islets by Free FA
The pancreatic islets are particularly susceptible to oxidative
stress. In the beginning of the 1980 decade, the activity of
antioxidant enzymes GPx, catalase, mitochondrial Mn-SOD
and cytosolic Cu/Zn-SOD was shown to be low in the endo-
crine pancreas when compared to other tissues.133 In 1996, the
Lenzen’s group reported that pancreatic islets have considerably
low mRNA levels of antioxidant enzymes in comparison with
the liver, kidney and brain. The islets present, in comparison
to the liver, about 15% of GPx, 38% of Cu/Zn-SOD, 30% of
Mn-SOD and very low expression of CAT 134 (Fig. 2). In the fol-
lowing year, this research group showed that the same was true
©2011 Landes Bioscience.
Do not distribute.
www.landesbioscience.com Islets 221
Acknowledgments
Thanks are due to Dr. Luiz R.G. Britto (University of São Paulo)
and Dr. Mauro Leonelli (University of São Paulo) for critically
reading the manuscript. Financial Support
Our studies have been supported by Fundação de Amparo à
Pesquisa do Estado de São Paulo (FAPESP), Conselho Nacional
de Desenvolvimento Cientifico e Tecnológico (CNPq), the
Instituto Nacional de Obesidade e Diabetes (INCT) and the US
Department of Veterans Affairs.
oxidative stress is another factor that contributes to tissue dys-
function caused by prolonged exposure to FA. In this review,
some therapeutic strategies to prevent β-cell dysfunction, such
as the use of antioxidant agents, modulation of UCP2 and
regulation of pro-oxidative signaling factors, such as NADPH
oxidase complex and RAS (both of which are common intra-
cellular signaling pathways modulated by FA), were indicated.
Nevertheless, more studies are undoubtedly necessary to clarify
the molecular mechanisms triggered by FA and involved in the
redox balance in pancreatic β-cells.
References
1. Pi J, Bai Y, Zhang Q, Wong V, Floering LM, Daniel K,
et al. Reactive oxygen species as a signal in glucose-stim-
ulated insulin secretion. Diabetes 2007; 56:1783-91.
2. Imoto H, Sasaki N, Iwase M, Nakamura U, Oku M,
Sonoki K, et al. Impaired insulin secretion by diphe-
nyleneiodium associated with perturbation of cytosolic
Ca2+ dynamics in pancreatic beta-cells. Endocrinology
2008; 149:5391-400.
3. Morgan D, Rebelato E, Abdulkader F, Graciano MF,
Oliveira-Emilio HR, Hirata AE, et al. Association of
NAD(P)H oxidase with glucose-induced insulin secre-
tion by pancreatic beta-cells. Endocrinology 2009;
150:2197-201.
4. Graciano MF, Santos LR, Curi R, Carpinelli AR.
NAD(P)H oxidase participates in the palmitate-
induced superoxide production and insulin secretion by
rat pancreatic islets. J Cell Physiol 2010; 226:1110-7.
5. Newsholme P, Haber EP, Hirabara SM, Rebelato EL,
Procopio J, Morgan D, et al. Diabetes associated cell
stress and dysfunction: role of mitochondrial and non-
mitochondrial ROS production and activity. J Physiol
(Lond) 2007; 583:9-24.
6. Baynes JW. Role of oxidative stress in development of
complications in diabetes. Diabetes 1991; 40:405-12.
7. Malaisse WJ, Malaisse-Lagae F, Wright PH. Effect of
fasting upon insulin secretion in the rat. Am J Physiol
1967; 213:843-8.
8. Boden G, Chen X, Iqbal N. Acute lowering of plasma
fatty acids lowers basal insulin secretion in diabetic and
nondiabetic subjects. Diabetes 1998; 47:1609-12.
9. Stein DT, Esser V, Stevenson BE, Lane KE, Whiteside
JH, Daniels MB, et al. Essentiality of circulating fatty
acids for glucose-stimulated insulin secretion in the
fasted rat. J Clin Invest 1996; 97:2728-35.
10. Carpinelli AR, Picinato MC, Stevanato E, Oliveira
HR, Curi R. Insulin secretion induced by palmitate—
a process fully dependent on glucose concentration.
Diabetes Metab 2002; 28:37-44.
11. Crespin SR, Greenough WB, 3rd, Steinberg D.
Stimulation of insulin secretion by infusion of free
fatty acids. J Clin Invest 1969; 48:1934-43.
12. Crespin SR, Greenough WB, Steinberg D. Stimulation
of insulin secretion by long-chain free fatty acids. A
direct pancreatic effect. J Clin Invest 1973; 52:1979-84.
13. Warnotte C, Gilon P, Nenquin M, Henquin JC.
Mechanisms of the stimulation of insulin release by
saturated fatty acids. A study of palmitate effects in
mouse beta-cells. Diabetes 1994; 43:703-11.
14. Poitout V, Hagman D, Stein R, Artner I, Robertson RP,
Harmon JS. Regulation of the insulin gene by glucose
and fatty acids. J Nutr 2006; 136:873-6.
15. Wang Y, Wang PY, Takashi K. Chronic effects of differ-
ent non-esterified fatty acids on pancreatic islets of rats.
Endocrine 2006; 29:169-73.
16. Zhou YP, Grill VE. Long-term exposure of rat pancre-
atic islets to fatty acids inhibits glucose-induced insulin
secretion and biosynthesis through a glucose fatty acid
cycle. J Clin Invest 1994; 93:870-6.
17. Prentki M, Joly E, El-Assaad W, Roduit R. Malonyl-
CoA signaling, lipid partitioning and glucolipotoxicity:
role in beta-cell adaptation and failure in the etiology
of diabetes. Diabetes 2002; 51:405-13.
18. Opara EC, Garfinkel M, Hubbard VS, Burch WM,
Akwari OE. Effect of fatty acids on insulin release:
role of chain length and degree of unsaturation. Am J
Physiol 1994; 266:635-9.
19. Stein DT, Stevenson BE, Chester MW, Basit M,
Daniels MB, Turley SD, et al. The insulinotropic
potency of fatty acids is influenced profoundly by their
chain length and degree of saturation. J Clin Invest
1997; 100:398-403.
20. Morgan NG, Dhayal S. Unsaturated fatty acids as
cytoprotective agents in the pancreatic beta-cell.
Prostaglandins Leukot Essent Fatty Acids 2010;
82:231-6.
21. Zhou YP, Grill VE. Palmitate-induced beta-cell insen-
sitivity to glucose is coupled to decreased pyruvate
dehydrogenase activity and enhanced kinase activity in
rat pancreatic islets. Diabetes 1995; 44:394-9.
22. Dobbins RL, Chester MW, Daniels MB, McGarry
JD, Stein DT. Circulating fatty acids are essential for
efficient glucose-stimulated insulin secretion after pro-
longed fasting in humans. Diabetes 1998; 47:1613-8.
23. Berne C. The metabolism of lipids in mouse pancreatic
islets. The oxidation of fatty acids and ketone bodies.
Biochem J 1975; 152:661-6.
24. Khan A, Ling ZC, Landau BR. Quantifying the car-
boxylation of pyruvate in pancreatic islets. J Biol Chem
1996; 271:2539-42.
25. MacDonald MJ. Glucose enters mitochondrial metab-
olism via both carboxylation and decarboxylation
of pyruvate in pancreatic islets. Metabolism 1993;
42:1229-31.
26. Sekine N, Cirulli V, Regazzi R, Brown LJ, Gine E,
Tamarit-Rodriguez J, et al. Low lactate dehydrogenase
and high mitochondrial glycerol phosphate dehydroge-
nase in pancreatic beta-cells. Potential role in nutrient
sensing. J Biol Chem 1994; 269:4895-902.
27. Farfari S, Schulz V, Corkey B, Prentki M. Glucose-
regulated anaplerosis and cataplerosis in pancreatic
beta-cells: possible implication of a pyruvate/citrate
shuttle in insulin secretion. Diabetes 2000; 49:718-26.
28. Deeney JT, Gromada J, Hoy M, Olsen HL, Rhodes
CJ, Prentki M, et al. Acute stimulation with long chain
acyl-CoA enhances exocytosis in insulin-secreting cells
(HIT T-15 and NMRI beta-cells). J Biol Chem 2000;
275:9363-8.
29. Prentki M, Vischer S, Glennon MC, Regazzi R, Deeney
JT, Corkey BE. Malonyl-CoA and long chain acyl-CoA
esters as metabolic coupling factors in nutrient-induced
insulin secretion. J Biol Chem 1992; 267:5802-10.
30. Prentki M, Tornheim K, Corkey BE. Signal transduc-
tion mechanisms in nutrient-induced insulin secretion.
Diabetologia 1997; 40:32-41.
31. Itoh Y, Kawamata Y, Harada M, Kobayashi M, Fujii
R, Fukusumi S, et al. Free fatty acids regulate insulin
secretion from pancreatic beta cells through GPR40.
Nature 2003; 422:173-6.
32. Kebede MA, Alquier T, Latour MG, Poitout V. Lipid
receptors and islet function: therapeutic implications?
Diabetes Obes Metab 2009; 11:10-20.
33. Briscoe CP, Tadayyon M, Andrews JL, Benson WG,
Chambers JK, Eilert MM, et al. The orphan G protein-
coupled receptor GPR40 is activated by medium and
long chain fatty acids. J Biol Chem 2003; 278:11303-11.
34. Kotarsky K, Nilsson NE, Olde B, Owman C. Progress
in methodology. Improved reporter gene assays used to
identify ligands acting on orphan seven-transmembrane
receptors. Pharmacol Toxicol 2003; 93:249-58.
35. Cartoni C, Yasumatsu K, Ohkuri T, Shigemura N,
Yoshida R, Godinot N, et al. Taste preference for fatty
acids is mediated by GPR40 and GPR120. J Neurosci
2010; 30:8376-82.
36. Edfalk S, Steneberg P, Edlund H. Gpr40 is expressed in
enteroendocrine cells and mediates free fatty acid stimu-
lation of incretin secretion. Diabetes 2008; 57:2280-7.
37. Parker HE, Habib AM, Rogers GJ, Gribble FM,
Reimann F. Nutrient-dependent secretion of glucose-
dependent insulinotropic polypeptide from primary
murine K cells. Diabetologia 2009; 52:289-98.
38. Latour MG, Alquier T, Oseid E, Tremblay C, Jetton
TL, Luo J, et al. GPR40 is necessary but not sufficient
for fatty acid stimulation of insulin secretion in vivo.
Diabetes 2007; 56:1087-94.
39. Lan H, Hoos LM, Liu L, Tetzloff G, Hu W,
Abbondanzo SJ, et al. Lack of FFAR1/GPR40 does
not protect mice from high-fat diet-induced metabolic
disease. Diabetes 2008; 57:2999-3006.
40. Ogawa T, Hirose H, Miyashita K, Saito I, Saruta T.
GPR40 gene Arg211His polymorphism may contrib-
ute to the variation of insulin secretory capacity in
Japanese men. Metabolism 2005; 54:296-9.
41. Vettor R, Granzotto M, De Stefani D, Trevellin E,
Rossato M, Farina MG, et al. Loss-of-function muta-
tion of the GPR40 gene associates with abnormal
stimulated insulin secretion by acting on intracellular
calcium mobilization. J Clin Endocrinol Metab 2008;
93:3541-50.
42. Martins EF, Miyasaka CK, Newsholme P, Curi R,
Carpinelli AR. Changes of fatty acid composition in
incubated rat pancreatic islets. Diabetes Metab 2004;
30:21-7.
43. Nolan CJ, Madiraju MS, Delghingaro-Augusto V,
Peyot ML, Prentki M. Fatty acid signaling in the {beta}-
cell and insulin secretion. Diabetes 2006; 55:16-23.
44. Pappan KL, Pan Z, Kwon G, Marshall CA, Coleman T,
Goldberg IJ, et al. Pancreatic beta-cell lipoprotein lipase
independently regulates islet glucose metabolism and
normal insulin secretion. J Biol Chem 2005; 280:9023-9.
45. Roduit R, Masiello P, Wang SP, Li H, Mitchell GA,
Prentki M. A role for hormone-sensitive lipase in
glucose-stimulated insulin secretion: a study in hor-
mone-sensitive lipase-deficient mice. Diabetes 2001;
50:1970-5.
46. Winzell MS, Svensson H, Arner P, Ahren B, Holm C.
The expression of hormone-sensitive lipase in clonal
beta-cells and rat islets is induced by long-term expo-
sure to high glucose. Diabetes 2001; 50:2225-30.
47. Roche E, Farfari S, Witters LA, Assimacopoulos-
Jeannet F, Thumelin S, Brun T, et al. Long-term expo-
sure of beta-INS cells to high glucose concentrations
increases anaplerosis, lipogenesis and lipogenic gene
expression. Diabetes 1998; 47:1086-94.
©2011 Landes Bioscience.
Do not distribute.
222 Islets Volume 3 Issue 5
85. Zhang CY, Baffy G, Perret P, Krauss S, Peroni O,
Grujic D, et al. Uncoupling protein-2 negatively
regulates insulin secretion and is a major link between
obesity, beta cell dysfunction and type 2 diabetes. Cell
2001; 105:745-55.
86. Rousset S, Mozo J, Dujardin G, Emre Y, Masscheleyn
S, Ricquier D, et al. UCP2 is a mitochondrial trans-
porter with an unusual very short half-life. FEBS Lett
2007; 581:479-82.
87. Schrauwen P, Saris WH, Hesselink MK. An alternative
function for human uncoupling protein 3: protection
of mitochondria against accumulation of nonesterified
fatty acids inside the mitochondrial matrix. FASEB J
2001; 15:2497-502.
88. Goglia F, Skulachev VP. A function for novel uncou-
pling proteins: antioxidant defense of mitochondrial
matrix by translocating fatty acid peroxides from the
inner to the outer membrane leaflet. FASEB J 2003;
17:1585-91.
89. Echtay KS, Roussel D, St-Pierre J, Jekabsons MB,
Cadenas S, Stuart JA, et al. Superoxide activates mito-
chondrial uncoupling proteins. Nature 2002; 415:96-9.
90. Echtay KS, Murphy MP, Smith RA, Talbot DA, Brand
MD. Superoxide activates mitochondrial uncoupling
protein 2 from the matrix side. Studies using targeted
antioxidants. J Biol Chem 2002; 277:47129-35.
91. Talbot DA, Duchamp C, Rey B, Hanuise N, Rouanet
JL, Sibille B, et al. Uncoupling protein and ATP/ADP
carrier increase mitochondrial proton conductance
after cold adaptation of king penguins. J Physiol 2004;
558:123-35.
92. Brand MD, Esteves TC. Physiological functions of the
mitochondrial uncoupling proteins UCP2 and UCP3.
Cell Metab 2005; 2:85-93.
93. Pi J, Bai Y, Daniel KW, Liu D, Lyght O, Edelstein D, et
al. Persistent oxidative stress due to absence of uncou-
pling protein 2 associated with impaired pancreatic
beta-cell function. Endocrinology 2009; 150:3040-8.
94. St-Pierre J, Buckingham JA, Roebuck SJ, Brand MD.
Topology of superoxide production from different sites
in the mitochondrial electron transport chain. J Biol
Chem 2002; 277:44784-90.
95. Koshkin V, Wang X, Scherer PE, Chan CB, Wheeler
MB. Mitochondrial functional state in clonal pancre-
atic beta-cells exposed to free fatty acids. J Biol Chem
2003; 278:19709-15.
96. Brand MD, Affourtit C, Esteves TC, Green K, Lambert
AJ, Miwa S, et al. Mitochondrial superoxide: produc-
tion, biological effects and activation of uncoupling
proteins. Free Radic Biol Med 2004; 37:755-67.
97. Joseph JW, Koshkin V, Zhang CY, Wang J, Lowell BB,
Chan CB, et al. Uncoupling protein 2 knockout mice
have enhanced insulin secretory capacity after a high-fat
diet. Diabetes 2002; 51:3211-9.
98. Ruzicka M, Skobisova E, Dlaskova A, Santorova J,
Smolkova K, Spacek T, et al. Recruitment of mito-
chondrial uncoupling protein UCP2 after lipopoly-
saccharide induction. Int J Biochem Cell Biol 2005;
37:809-21.
99. Medvedev AV, Robidoux J, Bai X, Cao W, Floering
LM, Daniel KW, et al. Regulation of the uncoupling
protein-2 gene in INS-1 beta-cells by oleic acid. J Biol
Chem 2002; 277:42639-44.
100. Surwit RS, Wang S, Petro AE, Sanchis D, Raimbault S,
Ricquier D, et al. Diet-induced changes in uncoupling
proteins in obesity-prone and obesity-resistant strains
of mice. Proc Natl Acad Sci USA 1998; 95:4061-5.
101. Sullivan PG, Rippy NA, Dorenbos K, Concepcion RC,
Agarwal AK, Rho JM. The ketogenic diet increases
mitochondrial uncoupling protein levels and activity.
Ann Neurol 2004; 55:576-80.
102. Sullivan PG, Dube C, Dorenbos K, Steward O, Baram
TZ. Mitochondrial uncoupling protein-2 protects the
immature brain from excitotoxic neuronal death. Ann
Neurol 2003; 53:711-7.
103. Bedard K, Krause KH. The NOX family of ROS-
generating NADPH oxidases: physiology and patho-
physiology. Physiol Rev 2007; 87:245-313.
65. Fridovich I. Superoxide anion radical (O2-.), superoxide
dismutases and related matters. J Biol Chem 1997;
272:18515-7.
66. Ceriello A, Motz E. Is oxidative stress the pathogenic
mechanism underlying insulin resistance, diabetes and
cardiovascular disease? The common soil hypothesis revis-
ited. Arterioscler Thromb Vasc Biol 2004; 24:816-23.
67. Griendling KK, Sorescu D, Ushio-Fukai M. NAD(P)H
oxidase: role in cardiovascular biology and disease. Circ
Res 2000; 86:494-501.
68. Victor VM, Apostolova N, Herance R, Hernandez-
Mijares A, Rocha M. Oxidative stress and mitochon-
drial dysfunction in atherosclerosis: mitochondria-
targeted antioxidants as potential therapy. Curr Med
Chem 2009; 16:4654-67.
69. Harraz MM, Marden JJ, Zhou W, Zhang Y, Williams
A, Sharov VS, et al. SOD1 mutations disrupt redox-
sensitive Rac regulation of NADPH oxidase in a famil-
ial ALS model. J Clin Invest 2008; 118:659-70.
70. Suvorava T, Kojda G. Reactive oxygen species as
cardiovascular mediators: lessons from endothelial-
specific protein overexpression mouse models. Biochim
Biophys Acta 2009; 1787:802-10.
71. Goldstein BJ, Mahadev K, Wu X. Redox paradox: insu-
lin action is facilitated by insulin-stimulated reactive
oxygen species with multiple potential signaling targets.
Diabetes 2005; 54:311-21.
72. Sivitz WI, Yorek MA. Mitochondrial dysfunction in
diabetes: from molecular mechanisms to functional
significance and therapeutic opportunities. Antioxid
Redox Signal 2010; 12:537-77.
73. Raha S, Robinson BH. Mitochondria, oxygen free
radicals, disease and ageing. Trends Biochem Sci 2000;
25:502-8.
74. Han D, Williams E, Cadenas E. Mitochondrial respira-
tory chain-dependent generation of superoxide anion
and its release into the intermembrane space. Biochem
J 2001; 353:411-6.
75. Morgan D, Oliveira-Emilio HR, Keane D, Hirata AE,
Santos da Rocha M, Bordin S, et al. Glucose, palmitate
and pro-inflammatory cytokines modulate production
and activity of a phagocyte-like NADPH oxidase in rat
pancreatic islets and a clonal beta cell line. Diabetologia
2007; 50:359-69.
76. Andreyev AY, Kushnareva YE, Starkov AA.
Mitochondrial metabolism of reactive oxygen species.
Biochemistry (Mosc) 2005; 70:200-14.
77. Bakker SJ, Ijzerman RG, Teerlink T, Westerhoff HV,
Gans RO, Heine RJ. Cytosolic triglycerides and oxida-
tive stress in central obesity: the missing link between
excessive atherosclerosis, endothelial dysfunction and
beta-cell failure? Atherosclerosis 2000; 148:17-21.
78. Loskovich MV, Grivennikova VG, Cecchini G,
Vinogradov AD. Inhibitory effect of palmitate on
the mitochondrial NADH:ubiquinone oxidoreductase
(complex I) as related to the active-de-active enzyme
transition. Biochem J 2005; 387:677-83.
79. Schonfeld P, Wojtczak L. Fatty acids as modulators of
the cellular production of reactive oxygen species. Free
Radic Biol Med 2008; 45:231-41.
80. Stewart JM, Blakely JA, Johnson MD. The interac-
tion of ferrocytochrome c with long-chain fatty acids
and their CoA and carnitine esters. Biochem Cell Biol
2000; 78:675-81.
81. Won JS, Singh I. Sphingolipid signaling and redox
regulation. Free Radic Biol Med 2006; 40:1875-88.
82. Wanders RJ, Waterham HR. Biochemistry of mamma-
lian peroxisomes revisited. Annu Rev Biochem 2006;
75:295-332.
83. Gehrmann W, Elsner M, Lenzen S. Role of metaboli-
cally generated reactive oxygen species for lipotoxicity
in pancreatic beta-cells. Diabetes Obes Metab 2010;
12:149-58.
84. Green K, Brand MD, Murphy MP. Prevention of mito-
chondrial oxidative damage as a therapeutic strategy in
diabetes. Diabetes 2004; 53:110-8.
48. Berthiaume L, Deichaite I, Peseckis S, Resh MD.
Regulation of enzymatic activity by active site fatty
acylation. A new role for long chain fatty acid acylation
of proteins. J Biol Chem 1994; 269:6498-505.
49. Yajima H, Komatsu M, Yamada S, Straub SG, Kaneko
T, Sato Y, et al. Cerulenin, an inhibitor of protein
acylation, selectively attenuates nutrient stimulation of
insulin release: a study in rat pancreatic islets. Diabetes
2000; 49:712-7.
50. Yamada S, Komatsu M, Sato Y, Yamauchi K, Aizawa
T, Kojima I. Nutrient modulation of palmitoylat-
ed 24-kilodalton protein in rat pancreatic islets.
Endocrinology 2003; 144:5232-41.
51. Komatsu M, Yajima H, Yamada S, Kaneko T, Sato Y,
Yamauchi K, et al. Augmentation of Ca2+-stimulated
insulin release by glucose and long-chain fatty acids
in rat pancreatic islets: free fatty acids mimic ATP-
sensitive K+ channel-independent insulinotropic action
of glucose. Diabetes 1999; 48:1543-9.
52. Fujiwara K, Maekawa F, Yada T. Oleic acid interacts
with GPR40 to induce Ca2+ signaling in rat islet beta-
cells: mediation by PLC and L-type Ca2+ channel and
link to insulin release. Am J Physiol Endocrinol Metab
2005; 289:670-7.
53. Shapiro H, Shachar S, Sekler I, Hershfinkel M, Walker
MD. Role of GPR40 in fatty acid action on the beta
cell line INS-1E. Biochem Biophys Res Commun
2005; 335:97-104.
54. Olofsson CS, Salehi A, Holm C, Rorsman P. Palmitate
increases L-type Ca2+ currents and the size of the readily
releasable granule pool in mouse pancreatic beta-cells. J
Physiol 2004; 557:935-48.
55. Su J, Yu H, Lenka N, Hescheler J, Ullrich S. The
expression and regulation of depolarization-activated
K+ channels in the insulin-secreting cell line INS-1.
Pflugers Arch 2001; 442:49-56.
56. Feng DD, Luo Z, Roh SG, Hernandez M, Tawadros
N, Keating DJ, et al. Reduction in voltage-gated K+
currents in primary cultured rat pancreatic beta-cells by
linoleic acids. Endocrinology 2006; 147:674-82.
57. Newsholme P, Morgan D, Rebelato E, Oliveira-Emilio
HC, Procopio J, Curi R, et al. Insights into the critical
role of NADPH oxidase(s) in the normal and dysregulat-
ed pancreatic beta cell. Diabetologia 2009; 52:2489-98.
58. Kaneto H, Katakami N, Kawamori D, Miyatsuka T,
Sakamoto K, Matsuoka TA, et al. Involvement of oxi-
dative stress in the pathogenesis of diabetes. Antioxid
Redox Signal 2007; 9:355-66.
59. Igoillo-Esteve M, Marselli L, Cunha DA, Ladriere
L, Ortis F, Grieco FA, et al. Palmitate induces a pro-
inflammatory response in human pancreatic islets
that mimics CCL2 expression by beta cells in type 2
diabetes. Diabetologia 53:1395-405.
60. Abaraviciene SM, Lundquist I, Salehi A. Rosiglitazone
counteracts palmitate-induced beta-cell dysfunction by
suppression of MAP kinase, inducible nitric oxide syn-
thase and caspase 3 activities. Cell Mol Life Sci 2008;
65:2256-65.
61. Maedler K, Spinas GA, Dyntar D, Moritz W, Kaiser N,
Donath MY. Distinct effects of saturated and monoun-
saturated fatty acids on beta-cell turnover and function.
Diabetes 2001; 50:69-76.
62. Karaskov E, Scott C, Zhang L, Teodoro T, Ravazzola
M, Volchuk A. Chronic palmitate but not oleate
exposure induces endoplasmic reticulum stress, which
may contribute to INS-1 pancreatic beta-cell apoptosis.
Endocrinology 2006; 147:3398-407.
63. Maestre I, Jordan J, Calvo S, Reig JA, Cena V, Soria B,
et al. Mitochondrial dysfunction is involved in apopto-
sis induced by serum withdrawal and fatty acids in the
beta-cell line INS-1. Endocrinology 2003; 144:335-45.
64. Kaneto H, Nakatani Y, Kawamori D, Miyatsuka T,
Matsuoka TA, Matsuhisa M, et al. Role of oxida-
tive stress, endoplasmic reticulum stress and c-Jun
N-terminal kinase in pancreatic beta-cell dysfunction
and insulin resistance. Int J Biochem Cell Biol 2005;
37:1595-608.
©2011 Landes Bioscience.
Do not distribute.
www.landesbioscience.com Islets 223
133. Grankvist K, Marklund SL, Taljedal IB. CuZn-
superoxide dismutase, Mn-superoxide dismutase, cata-
lase and glutathione peroxidase in pancreatic islets and
other tissues in the mouse. Biochem J 1981; 199:393-8.
134. Lenzen S, Drinkgern J, Tiedge M. Low antioxidant
enzyme gene expression in pancreatic islets compared
with various other mouse tissues. Free Radic Biol Med
1996; 20:463-6.
135. Tiedge M, Lortz S, Drinkgern J, Lenzen S. Relation
between antioxidant enzyme gene expression and anti-
oxidative defense status of insulin-producing cells.
Diabetes 1997; 46:1733-42.
136. Tran PO, Parker SM, LeRoy E, Franklin CC, Kavanagh
TJ, Zhang T, et al. Adenoviral overexpression of the
glutamylcysteine ligase catalytic subunit protects pan-
creatic islets against oxidative stress. J Biol Chem 2004;
279:53988-93.
137. Rashidi A, Kirkwood TB, Shanley DP. Metabolic
evolution suggests an explanation for the weakness of
antioxidant defences in beta-cells. Mech Ageing Dev
2009; 130:216-21.
138. Tonooka N, Oseid E, Zhou H, Harmon JS, Robertson
RP. Glutathione peroxidase protein expression and
activity in human islets isolated for transplantation.
Clin Transplant 2007; 21:767-72.
139. Cornelius JG, Luttge BG, Peck AB. Antioxidant
enzyme activities in IDD-prone and IDD-resistant
mice: a comparative study. Free Radic Biol Med 1993;
14:409-20.
140. Ladyman SR, Augustine RA, Grattan DR. Hormone
interactions regulating energy balance during preg-
nancy. J Neuroendocrinol 2010; 22:805-17.
141. Fehm HL, Kern W, Peters A. The selfish brain: com-
petition for energy resources. Prog Brain Res 2006;
153:129-40.
142. Rebelato E, Abdulkader F, Curi R, Carpinelli AR. Low
doses of hydrogen peroxide impair glucose-stimulated
insulin secretion via inhibition of glucose metabo-
lism and intracellular calcium oscillations. Metabolism
2010; 59:409-13.
143. Vandewalle B, Moerman E, Lefebvre B, Defrance F,
Gmyr V, Lukowiak B, et al. PPARgamma-dependent
and -independent effects of rosiglitazone on lipo-
toxic human pancreatic islets. Biochem Biophys Res
Commun 2008; 366:1096-101.
144. Bikopoulos G, da Silva Pimenta A, Lee SC, Lakey
JR, Der SD, Chan CB, et al. Ex vivo transcriptional
profiling of human pancreatic islets following chronic
exposure to monounsaturated fatty acids. J Endocrinol
2008; 196:455-64.
119. Tahmasebi M, Puddefoot JR, Inwang ER, Vinson GP.
The tissue renin-angiotensin system in human pan-
creas. J Endocrinol 1999; 161:317-22.
120. Carlsson PO, Berne C, Jansson L. Angiotensin II and
the endocrine pancreas: effects on islet blood flow and
insulin secretion in rats. Diabetologia 1998; 41:127-33.
121. Hirata AE, Morgan D, Oliveira-Emilio HR, Rocha
MS, Carvalho CR, Curi R, et al. Angiotensin II induces
superoxide generation via NAD(P)H oxidase activation
in isolated rat pancreatic islets. Regul Pept 2009; 153:1-6.
122. Lupi R, Del Guerra S, Bugliani M, Boggi U, Mosca F,
Torri S, et al. The direct effects of the angiotensin-con-
verting enzyme inhibitors, zofenoprilat and enalaprilat,
on isolated human pancreatic islets. Eur J Endocrinol
2006; 154:355-61.
123. Chu KY, Lau T, Carlsson PO, Leung PS. Angiotensin
II type 1 receptor blockade improves beta-cell function
and glucose tolerance in a mouse model of type 2 dia-
betes. Diabetes 2006; 55:367-74.
124. Nakayama M, Inoguchi T, Sonta T, Maeda Y, Sasaki
S, Sawada F, et al. Increased expression of NAD(P)H
oxidase in islets of animal models of Type 2 diabetes
and its improvement by an AT1 receptor antagonist.
Biochem Biophys Res Commun 2005; 332:927-33.
125. Vitale C, Mercuro G, Castiglioni C, Cornoldi A, Tulli
A, Fini M, et al. Metabolic effect of telmisartan and
losartan in hypertensive patients with metabolic syn-
drome. Cardiovasc Diabetol 2005; 4:1-8.
126. Benson SC, Pershadsingh HA, Ho CI, Chittiboyina
A, Desai P, Pravenec M, et al. Identification of telmis-
artan as a unique angiotensin II receptor antago-
nist with selective PPARgamma-modulating activity.
Hypertension 2004; 43:993-1002.
127. Tikellis C, Wookey PJ, Candido R, Andrikopoulos S,
Thomas MC, Cooper ME. Improved islet morphology
after blockade of the rennin-angiotensin system in the
ZDF rat. Diabetes 2004; 53:989-97.
128. Saitoh Y, Hongwei W, Ueno H, Mizuta M, Nakazato
M. Telmisartan attenuates fatty-acid-induced oxidative
stress and NAD(P)H oxidase activity in pancreatic
beta-cells. Diabetes Metab 2009; 35:392-7.
129. Saitoh Y, Hongwei W, Ueno H, Mizuta M, Nakazato
M. Candesartan attenuates fatty acid-induced oxidative
stress and NAD(P)H oxidase activity in pancreatic
beta-cells. Diabetes Res Clin Pract 2010; 90:54-9.
130. Veluthakal R, Madathilparambil SV, McDonald P,
Olson LK, Kowluru A. Regulatory roles for Tiam1,
a guanine nucleotide exchange factor for Rac1, in
glucose-stimulated insulin secretion in pancreatic beta-
cells. Biochem Pharmacol 2009; 77:101-13.
131. Mertens AE, Roovers RC, Collard JG. Regulation of
Tiam1-Rac signalling. FEBS Lett 2003; 546:11-6.
132. Syed I, Jayaram B, Subasinghe W, Kowluru A. Tiam1/
Rac1 signaling pathway mediates palmitate-induced,
ceramide-sensitive generation of superoxides and lipid
peroxides and the loss of mitochondrial membrane
potential in pancreatic beta-cells. Biochem Pharmacol
2010; 80:874-83.
104. Oliveira HR, Verlengia R, Carvalho CR, Britto LR,
Curi R, Carpinelli AR. Pancreatic beta-cells express
phagocyte-like NAD(P)H oxidase. Diabetes 2003;
52:1457-63.
105. Uchizono Y, Takeya R, Iwase M, Sasaki N, Oku M,
Imoto H, et al. Expression of isoforms of NADPH
oxidase components in rat pancreatic islets. Life Sci
2006; 80:133-9.
106. Vilhardt F, van Deurs B. The phagocyte NADPH
oxidase depends on cholesterol-enriched membrane
microdomains for assembly. EMBO J 2004; 23:739-48.
107. Zhang AY, Yi F, Zhang G, Gulbins E, Li PL. Lipid raft
clustering and redox signaling platform formation in
coronary arterial endothelial cells. Hypertension 2006;
47:74-80.
108. Hilenski LL, Clempus RE, Quinn MT, Lambeth
JD, Griendling KK. Distinct subcellular localizations
of Nox1 and Nox4 in vascular smooth muscle cells.
Arterioscler Thromb Vasc Biol 2004; 24:677-83.
109. Cohen AW, Razani B, Schubert W, Williams TM,
Wang XB, Iyengar P, et al. Role of caveolin-1 in the
modulation of lipolysis and lipid droplet formation.
Diabetes 2004; 53:1261-70.
110. Fisher AB. Redox signaling across cell membranes.
Antioxid Redox Signal 2009; 11:1349-56.
111. Valle MM, Graciano MF, Lopes de Oliveira ER,
Camporez JP, Akamine EH, Carvalho CR, et al.
Alterations of NADPH oxidase activity in rat pancre-
atic islets induced by a high-fat diet. Pancreas 2011;
40:390-5.
112. Mikami A, Imoto K, Tanabe T, Niidome T, Mori Y,
Takeshima H, et al. Primary structure and functional
expression of the cardiac dihydropyridine-sensitive
calcium channel. Nature 1989; 340:230-3.
113. Yuan H, Zhang X, Huang X, Lu Y, Tang W, Man Y,
et al. NADPH oxidase 2-derived reactive oxygen spe-
cies mediate FFAs-induced dysfunction and apoptosis
of beta-cells via JNK, p38 MAPK and p53 pathways.
PLoS One 5:15726.
114. Michalska M, Wolf G, Walther R, Newsholme P.
Effects of pharmacological inhibition of NADPH oxi-
dase or iNOS on pro-inflammatory cytokine, palmitic
acid or H2O2-induced mouse islet or clonal pancreatic
beta-cell dysfunction. Biosci Rep 2010; 30:445-53.
115. Keane DC, Takahashi HK, Dhayal S, Morgan NG,
Curi R, Newsholme P. Arachidonic acid actions on
functional integrity and attenuation of the negative
effects of palmitic acid in a clonal pancreatic beta-cell
line. Clin Sci (Lond) 2011; 120:195-206.
116. Lau T, Carlsson PO, Leung PS. Evidence for a local
angiotensin-generating system and dose-dependent
inhibition of glucose-stimulated insulin release by
angiotensin II in isolated pancreatic islets. Diabetologia
2004; 47:240-8.
117. Leung PS, Chappell MC. A local pancreatic renin-
angiotensin system: endocrine and exocrine roles. Int J
Biochem Cell Biol 2003; 35:838-46.
118. Leung PS, Carlsson PO. Tissue renin-angiotensin
system: its expression, localization, regulation and
potential role in the pancreas. J Mol Endocrinol 2001;
26:155-64.
... Metabolism of sorbitol, hexosamine, or methylglyoxal derived from glucose catabolism can also stimulate ROS production in diabetic conditions [8,18]. However, mitochondrial ROS production in β-cells has been documented mainly under conditions of the increased flux of fatty acids and amino acids as substrates, the conditions associated with diabetic pathology [11,[19][20][21][22][23][24] (Fig. 1). Superoxide produced along the electron transport chain (complex I/III) is rapidly converted by superoxide dismutase, SOD2 isoform, to H2O2, being able to diffuse out of mitochondria [19,[24][25][26][27][28][29] This production has been found to be controlled by the uncoupling activity of UCP2 in β-cells [30][31][32]. ...
Article
Redox status plays a multifaceted role in the intricate physiology and pathology of pancreatic beta-cells, the pivotal regulators of glucose homeostasis through insulin secretion. They are highly responsive to changes in metabolic cues where reactive oxygen species are part of it, all arising from nutritional intake. These molecules not only serve as crucial signaling intermediates for insulin secretion but also participate in the nuanced heterogeneity observed within the beta-cell population. A central aspect of beta-cell redox biology revolves around the localized production of hydrogen peroxide and the activity of NADPH oxidases which are tightly regulated and serve diverse physiological functions. Pancreatic beta-cells possess a remarkable array of antioxidant defense mechanisms although considered relatively modest compared to other cell types, are efficient in preserving redox balance within the cellular milieu. This intrinsic antioxidant machinery operates in concert with redox-sensitive signaling pathways, forming an elaborate redox relay system essential for beta-cell function and adaptation to changing metabolic demands. Perturbations in redox homeostasis can lead to oxidative stress exacerbating insulin secretion defect being a hallmark of type 2 diabetes. Understanding the interplay between redox signaling, oxidative stress, and beta-cell dysfunction is paramount for developing effective therapeutic strategies aimed at preserving beta-cell health and function in individuals with type 2 diabetes. Thus, unraveling the intricate complexities of beta-cell redox biology presents exciting avenues for advancing our understanding and treatment of metabolic disorders.
... Pancreatic islets, unlike the liver that actively detoxifies xenobiotics, do not possess extensive antioxidant defense machinery. Additionally, ROS in β-cells play an important role in cellular signaling and insulin secretion [53,54], thus β-cells are more sensitive to redox modulation by DMF. ...
Article
Dimethyl fumarate (DMF) is pharmaceutical activator of the transcription factor nuclear factor erythroid 2-related factor 2 (Nrf2), which regulates of many cellular antioxidant response pathways, and has been used to treat inflammatory diseases such as multiple sclerosis. However, DMF has been shown to produce adverse effects on offspring in animal studies and as such is not recommended for use during pregnancy. The goal of this work is to better understand how these adverse effects are initiated and the role of DMF-induced Nrf2 activation during three critical windows of development in embryonic zebrafish (Danio rerio): pharyngula, hatching, and protruding-mouth stages. To evaluate Nrf2 activation, wildtype zebrafish, and mutant zebrafish (nrf2afh318/fh318) embryos with a loss of function mutation in Nrf2a, the co-ortholog to human Nrf2, were treated for 6 h with DMF (0-20 μM) beginning at the pharyngula, hatching, or protruding-mouth stage and assessed for survival and morphology. Nrf2a mutant fish had an increase in survival, however, morphology studies demonstrated Nrf2a mutant fish had more severe deformities occurring with exposures during the hatching stage. To verify Nrf2 cellular localization and downstream impacts on protein-S-glutathionylation in situ, a concentration below the LOAEL was chosen (7 μM) for immunohistochemistry and S-glutathionylation. Embryos were imaged via epifluorescence microscopy studies, the Nrf2a protein in the body tissue was decreased with DMF only when exposed at the hatching stage, while total protein S-glutathionylation was modulated by Nrf2a activity and DMF during the pharyngula and protruding-mouth stage. The pancreatic islet and liver were further analyzed via confocal microscopy. Pancreatic islets and liver also had tissue specific differences with Nrf2a protein expression and protein S-glutathionylation. This work demonstrates how critical windows of exposure and Nrf2a activity may influence toxicity of DMF and highlights tissue-specific changes in Nrf2a protein levels and S-glutathionylation in pancreatic islet and liver during embryonic development.
... Additionally, we assessed the function of post-cultured islets by measuring the glucosestimulated insulin secretion (figure 9(C)). Islets cultured in CTL secreted high insulin levels even in the low glucose environment and did not respond to the high glucose, indicating the islet damage [54,55], In contrast, islets cultured in micropyramid + O 2 showed the well-regulated insulin secretion in the low glucose environment and responded well to the high glucose solution. Islets cultured in micropyramid condition secreted insulin levels between the other two conditions. ...
Article
Full-text available
The need for maintaining cell-spheroid viability and function within high-density cultures is unmet for various clinical and experimental applications, including cell therapies. One immediate application is for transplantation of pancreatic islets, a clinically recognized treatment option to cure type 1 diabetes; islets are isolated from a donor for subsequent culture prior to transplantation. However, high seeding conditions cause unsolicited fusion of multiple spheroids, thereby limiting oxygen diffusion to induce hypoxic cell death. Here we introduce a culture dish incorporating a Micropyramid-patterned surface to prevent the unsolicited fusion and oxygen-permeable bottom for optimal oxygen environment. A 400 µm-thick, oxygen-permeable polydimethylsiloxane sheet topped with Micropyramid pattern of 400 µm-base and 200 µm-height was fabricated to apply to the 24-well plate format. The Micropyramid pattern separated the individual pancreatic islets to prevent the fusion of multiple islets. This platform supported the high oxygen demand of islets at high seeding density at 260 islet equivalents/cm2, a 2 – 3-fold higher seeding density compared to the conventional islet culture used in a preparation for the clinical islet transplantations, demonstrating improved islet morphology, metabolism and function in a 4 day-culture. Transplantation of these islets into immunodeficient diabetic mice exhibited significantly improved engraftment to achieve euglycemia compared to islets cultured in the conventional culture wells. Collectively, this simple design modification allows for high-density cultures of three-dimensional cell spheroids to improve the viability and function for an array of investigational and clinical replacement tissues.
Article
Both high glucose intake with a high-fat meal and inhibition of dipeptidyl peptidase-4 (DPP4) have been associated with plasma lipid-lowering effects, but mechanistic understanding linking glucose and fat absorption is lacking. We here hypothesized that glucose ameliorates intestinal fatty acid uptake via a pathway involving DPP4. A concentration of 50 mM glucose reduced mean DPP4 activity in differentiated Caco-2 enterocytes by 42.5 % and fatty acid uptake by 66.0 % via nutrient sensing by the sweet taste receptor subunit TAS1R3 and glucose transporter GLUT-2. No effect of the DPP4 substrates GLP-1 and GIP or of the cellular energy status on the reduced uptake of fatty acids was seen, but a direct interaction between DPP4 and fatty acid transporters is suggested. Conclusively we identified DPP4 as a regulator of fatty acid absorption in Caco-2 enterocytes that mediates the inhibition of intestinal fatty acid uptake by glucose via an interplay of GLUT-2 and TAS1R3.
Article
Full-text available
In Type 1 and Type 2 diabetes, pancreatic β-cell survival and function are impaired. Additional etiologies of diabetes include dysfunction in insulin-sensing hepatic, muscle, and adipose tissues as well as immune cells. An important determinant of metabolic health across these various tissues is mitochondria function and structure. This review focuses on the role of mitochondria in diabetes pathogenesis, with a specific emphasis on pancreatic β-cells. These dynamic organelles are obligate for β-cell survival, function, replication, insulin production, and control over insulin release. Therefore, it is not surprising that mitochondria are severely defective in diabetic contexts. Mitochondrial dysfunction poses challenges to assess in cause-effect studies, prompting us to assemble and deliberate the evidence for mitochondria dysfunction as a cause or consequence of diabetes. Understanding the precise molecular mechanisms underlying mitochondrial dysfunction in diabetes and identifying therapeutic strategies to restore mitochondrial homeostasis and enhance β-cell function are active and expanding areas of research. In summary, this review examines the multidimensional role of mitochondria in diabetes, focusing on pancreatic β-cells and highlighting the significance of mitochondrial metabolism, bioenergetics, calcium, dynamics, and mitophagy in the pathophysiology of diabetes. We describe the effects of diabetes-related gluco/lipotoxic, oxidative and inflammation stress on β-cell mitochondria, as well as the role played by mitochondria on the pathologic outcomes of these stress paradigms. By examining these aspects, we provide updated insights and highlight areas where further research is required for a deeper molecular understanding of the role of mitochondria in β-cells and diabetes.
Article
Full-text available
Type 2 Diabetes Mellitus (T2DM) has been defined as one of the commonly occurring metabolic disorders arising due to the combination of 2 primary factors, namely, inadequate insulin response in the insulin-sensitive tissues and insufficient insulin secretion through pancreatic beta-cells. The precise regulation of insulin release, detection, and synthesis is imperative for glucose homeostasis maintenance, and involves intricate molecular mechanisms. The onset of the disease may be attributed to a metabolic imbalance arising from deficiencies in any of the mechanisms in processes. This research examines the primary characteristics of T2DM, including the molecular mechanisms and pathways contributing to the insulin resistance (IR) and T2DMdevelopment. The information is compiled with the aim of examining insulin release, sensing, synthesis, and resulting effects on various insulin-sensitive organs. Special attention is paid to these aspects. This paper examines the various pathogenic factors contributing to T2DM development and progression, including diet, exercise, gut dysbiosis, and metabolic memory. The present discourse delves into the molecular pathways which establish a connection between IR and T2DM. Furthermore, the discussion highlights the noteworthy implications of T2DM, particularly the hastened progression of atherosclerosis, on cardiovascular risk.
Article
Background Uncoupling proteins (UCPs) are unpaired electron carriers that uncouple oxygen intake by the electron transport chain from ATP production in the inner membrane of the mitochondria. The physiological activities of UCPs have been hotly contested, and the involvement of UCPs in the pathogenesis and progression of diabetes mellitus is among the greatest concerns. UCPs are hypothesised to be triggered by superoxide and then reduce mitochondrial free radical production, potentially protecting diabetes mellitus patients who are experiencing oxidative stress. Objectives The objectives of the study are to find out the newest ways to treat diabetes mellitus through protein uncoupling. Methods Research and review papers are collected from different databases like google scholar, PubMed, Mendeley, Scopus, Science Open, Directory of open access journals, and Education Resources Information Center, using different keywords such as “uncoupling proteins in diabetes mellitus treatment”, “UCP 1”, “UCP 2”, and ‘UCP 3”. Results UCP1, UCP2, and UCP 3 are the potential targets as an uncoupling protein for the treatment of diabetes mellitus for new drugs. New drugs treat the disease by reducing oxidative stress through thermogenesis and energy expenditure. Conclusion UCP1, UCP2, and UCP3 have a role in fatty acid metabolism, negative control of insulin production, and insulin sensitivity by beta-cells. Polymorphisms in the UCP 1, 2, and 3 genes significantly reduce the risk of developing diabetes mellitus. Protein uncoupling indirectly targets the GPCR and islet of Langerhans. This review summarises the advances in understanding the role of UCP1, UCP2, and UCP3 in diabetes mellitus.
Article
Full-text available
Fatty acids (FAs) and other lipid molecules are important for many cellular functions, including vesicle exocytosis. For the pancreatic beta-cell, while the presence of some FAs is essential for glucose-stimulated insulin secretion, FAs have enormous capacity to amplify glucose-stimulated insulin secretion, which is particularly operative in situations of D-cell compensation for insulin resistance. In this review, we propose that FAs do this via three interdependent processes, which we have assigned to a "trident model" of P-cell lipid signaling. The first two arms of the model implicate intracellular metabolism of FAs, whereas the third is related to membrane free fatty acid receptor (FFAR) activation. The first arm involves the AMP-activated protein kinase/malonyl-CoA/long-chain acyl-CoA (LC-CoA) signaling network in which glucose, together with other anaplerotic fuels, increases cytosolic malonyl-CoA, which inhibits FA partitioning into oxidation, thus increasing the availability of LC-CoA for signaling purposes. The second involves glucose-responsive triglyceride (TG)/free fatty acid (FFA) cycling. In this pathway, glucose promotes LC-CoA esterification to complex lipids such as TG and diacylglycerol, concomitant with glucose stimulation of lipolysis of the esterification products, with renewal of the intracellular FFA pool for reactivation to LC-CoA. The third arm involves FFA stimulation of the G-protein-coupled receptor GPR40/FFAR1, which results in enhancement of glucose-stimulated accumulation of cytosolic Ca2+ and consequently insulin secretion. It is possible that FFA released by the lipolysis arm of TG/FFA cycling is partly "secreted" and, via an autocrine/paracrine mechanism, is additive to exogenous FFAs in activating the FFAR1 pathway. Glucose-stimulated release of arachidonic acid from phospholipids by calcium-independent phospholipase A(2) and/or from TG/FFA cycling may also be involved. Improved knowledge of lipid signaling in the beta-cell will allow a better understanding of the mechanisms of beta-cell compensation and failure in diabetes.
Article
Full-text available
Uncoupling protein 2 (UCP2) maps to a region on distal mouse chromosome 7 that has been linked to the phenotypes of obesity and type II diabetes. We recently reported that UCP2 expression is increased by high fat feeding in adipose tissue of the A/J strain of mice, which is resistant to the development of dietary obesity. More recently, a third UCP (UCP3) was identified, which is expressed largely in skeletal muscle and brown adipose tissue. The UCP2 and UCP3 genes are located adjacent to one another on mouse chromosome 7. Thus, the roles of these UCPs in both metabolic efficiency and the linkage to obesity and diabetes syndromes is unclear. For this reason, we examined the expression of UCP2 and UCP3 in white adipose tissue and interscapular brown adipose tissue and in gastrocnemius/soleus muscle preparations from the obesity-resistant A/J and C57BL/KsJ (KsJ) strains and the obesity-prone C57BL/6J (B6) mouse strain. In both KsJ and A/J mice, UCP2 expression in white fat was increased ≈2-fold in response to 2 weeks of a high fat diet, but there was no effect of diet on UCP2 levels in B6 mice. In skeletal muscle and in brown fat, neither UCP2 nor UCP3 expression was affected by diet in A/J, B6, or KsJ mice. However, in brown fat, we observed a 2–3-fold increase in the expression of UCP1 in response to dietary fat challenge, which may be related to diet-induced elevations in plasma leptin levels. Together, these results indicate that the consumption of a high fat diet selectively regulates UCP2 expression in white fat and UCP1 expression in brown fat and that resistance to obesity is correlated with this early, selective induction of UCP1 and UCP2 and is not associated with changes in expression of UCP3.
Article
Full-text available
Non-insulin-dependent diabetes mellitus is associated with, in addition to impaired insulin release, elevated levels of free fatty acids (FFA) in the blood. Insulin release is stimulated when β-cells are acutely exposed to FFA, whereas chronic exposure may inhibit glucose-induced insulin secretion. In the present study we investigated the direct effects of long chain acyl-CoA (LC-CoA), the active intracellular form of FFA, on insulin exocytosis. Palmitoyl-CoA stimulated both insulin release from streptolysin-O-permeabilized HIT cells and fusion of secretory granules to the plasma membrane of mouse pancreatic β-cells, as measured by cell capacitance. The LC-CoA effect was chain length-dependent, requiring chain lengths of at least 14 carbons. LC-CoA needed to be present to stimulate insulin release, and consequently there was no effect following its removal. The stimulatory effect was observed after inhibition of protein kinase activity and in the absence of ATP, even though both kinases and ATP, themselves, modulate exocytosis. The effect of LC-CoA was inhibited by cerulenin, which has been shown to block protein acylation. The data suggest that altered LC-CoA levels, resulting from FFA or glucose metabolism, may act directly on the exocytotic machinery to stimulate insulin release by a mechanism involving LC-CoA protein binding.
Article
Antioxidant enzyme expression was determined in rat pancreatic islets and RINm5F insulin-producing cells on the level of mRNA, protein, and enzyme activity in comparison with 11 other rat tissues. Although superoxide dismutase expression was in the range of 30% of the liver values, the expression of the hydrogen peroxide-inactivating enzymes catalase and glutathione per-oxidase was extremely low, in the range of 5% of the liver. Pancreatic islets but not RINm5F cells expressed an additional phospholipid hydroperoxide glutathione peroxidase that exerted protective effects against lipid peroxidation of the plasma membrane. Regression analysis for mRNA and protein expression and enzyme activities from 12 rat tissues revealed that the mRNA levels determine the enzyme activities of the tissues. The induction of cellular stress by high glucose, high oxygen, and heat shock treatment did not affect antioxidant enzyme expression in rat pancreatic islets or in RINm5F cells. Thus insulin-producing cells cannot adapt the low antioxidant enzyme activity levels to typical situations of cellular stress by an upregulation of gene expression. Through stable transfection, however, we were able to increase catalase and glutathione peroxidase gene expression in RINm5F cells, resulting in enzyme activities more than 100-fold higher than in nontransfected controls. Catalase-transfected RINm5F cells showed a 10-fold greater resistance toward hydrogen peroxide toxicity, whereas glutathione peroxidase overexpression was much less effective. Thus inactiva-tion of hydrogen peroxide through catalase seems to be a step of critical importance for the removal of reactive oxygen species in insulin-producing cells. Overexpression of catalase may therefore be an effective means of preventing the toxic action of reactive oxygen species.
Article
The rate of oxidation of 14C-labelled fatty acids and of ketone bodies was measured in isolated pancreatic islets of obese-hyperglycaemic mice (ob/ob). The following main observations were made. 1. Octanoate, palmitate and oleate were all converted into CO2 by the pancreatic islets. Octanoate was oxidized with the highest rate followed by palmitate and oleate. 2. The rate of oxidation of 0.7 mM-palmitate was 3.1 pmol/h per mug drug weight. This was decreased by 50% in the presence of 16.7 mM-glucose. The rate of palmitate oxidation was also inhibited by 2-bromostearate. The palmitate oxidation showed a concentration-dependent increase, which was most marked between 0.25 and 1.0 mM. 3. Octanoate (5 mM) had no effect on the rate of oxidation of 3.3 mM- glucose. 4. Acetoacetate (5 mM) and D-3-hydroxybutyrate (5 mM) were oxidized at rates of 5.9 and 5.4 pmol/h per mug dry weight respectively. These rates were less than 10% of those found in kidney-cortex slices. The magnitude of the oxidation rates found for fatty acids and for ketone bodies suggest that these substrates represent important metabolic fuels for the pancreatic B-cells.
Article
Oxidative stress is considered a major contributor to etiology of both "normal" senescence and severe pathologies with serious public health implications. Mitochondria generate reactive oxygen species (ROS) that are thought to augment intracellular oxidative stress. Mitochondria possess at least nine known sites that are capable of generating superoxide anion, a progenitor ROS. Mitochondria also possess numerous ROS defense systems that are much less studied. Studies of the last three decades shed light on many important mechanistic details of mitochondrial ROS production, but the bigger picture remains obscure. This review summarizes the current knowledge about major components involved in mitochondrial ROS metabolism and factors that regulate ROS generation and removal. An integrative, systemic approach is applied to analysis of mitochondrial ROS metabolism, which is now dissected into mitochondrial ROS production, mitochondrial ROS removal, and mitochondrial ROS emission. It is suggested that mitochondria augment intracellular oxidative stress due primarily to failure of their ROS removal systems, whereas the role of mitochondrial ROS emission is yet to be determined and a net increase in mitochondrial ROS production in situ remains to be demonstrated.
Article
Evidence exists for the presence of a discrete tissue renin-angiotensin system (RAS) in mouse and rat pancreas that is thought largely to be associated with the vasculature. To investigate this in the human pancreas, and to establish whether the cellular sites of RAS components include the islets of Langerhans, we used immunocytochemistry to localise the expression of angiotensin II (AT1) receptors and (pro)renin, and non-isotopic in situ hybridisation to localise transcription of the (pro)renin gene. Identification of cell types in the islets of Langerhans was achieved using antibodies to glucagon and insulin. The results show the presence of the AT1 receptor and (pro)renin both in the beta cells of the islets of Langerhans, and in endothelial cells of the pancreatic vasculature. Transcription of (pro)renin mRNA, however, was confined to connective tissue surrounding the blood vessels and in reticular fibres within the islets. These findings are similar to those obtained in other tissues, and suggest that renin may be released from its sites of synthesis and taken up by possible cellular sites of action. The results presented here suggest that a tissue RAS may be present in human pancreas and that it may directly affect beta cell function as well as pancreatic blood flow.
Article
The aim of the present study was to characterize depolarization-activated outward currents in insulin-secreting INS-1 cells and to investigate the role of K+ channels other than the KATP channels in the regulation of insulin release. Outward currents were inhibited by 4-aminopyridine (4-AP, 10 mmol/l), tetraethylammonium (TEA, 10 mmol/l) and tetrapentylammonium (TPeA, 100 mol/l) by 55.1&#453.8% (n=3), 78.1&#453.2% (n=6) and 98.7&#450.8% (n=5), respectively. Margatoxin (5 nmol/l) and charybdotoxin (3 mol/l) had no effect. 4-AP inhibited mainly a fast-activating, slowly inactivating current, whereas the TEA- and TPeA-sensitive current components were slowly activating and non-inactivating. Forskolin and the forskolin analogue 1,9-dideoxyforskolin, which does not stimulate adenylyl cyclase, also inhibited the outward current, suggesting a direct effect on the channels. Using reverse transcriptase polymerase chain reaction (RT/PCR), Kv channel mRNAs of Kv1.4, Kv1.5, Kv2.1, Kv2.2, Kv3.1 and Kv3.2 were detected whereas other Kv channels, Kv1.1, Kv1.2, Kv1.3, Kv1.6 and Kv3.4 were not detected. Insulin secretion in the presence of tolbutamide (100 mol/l) was increased by 4-AP, TEA and TPeA by 65%, 41% and 150%, respectively. Basal secretion was not affected by these blockers. Our study reveals that the opening of voltage-dependent K+ channels negatively controls insulin secretion in depolarized cells, probably by shortening the action potential thus reducing Ca2+ influx.