Open access peer-reviewed chapter

Redox Signaling is Essential for Insulin Secretion

Written By

Petr Ježek, Blanka Holendová, Martin Jabůrek, Jan Tauber, Andrea Dlasková and Lydie Plecitá-Hlavatá

Submitted: 20 June 2020 Reviewed: 02 October 2020 Published: 18 June 2021

DOI: 10.5772/intechopen.94312

From the Edited Volume

Type 2 Diabetes - From Pathophysiology to Cyber Systems

Edited by Anca Pantea Stoian

Chapter metrics overview

376 Chapter Downloads

View Full Metrics

Abstract

In this review, we place redox signaling in pancreatic β-cells to the context with signaling pathways leading to insulin secretion, acting for example upon the action of incretins (GLP-1, GIP) and the metabotropic receptor GPR40. Besides a brief description of ion channel participation in depolarization/repolarization of the plasma membrane, we emphasize a prominent role of the elevated glucose level in pancreatic β-cells during glucose-stimulated insulin secretion (GSIS). We focus on our recent findings, which revealed that for GSIS, not only elevated ATP synthesis is required, but also fundamental redox signaling originating from the NADPH oxidase 4- (NOX4-) mediated H2O2 production. We hypothesized that the closing of the ATP-sensitive K+ channel (KATP) is only possible when both ATP plus H2O2 are elevated in INS-1E cells. KATP alone or with synergic channels provides an element of logical sum, integrating both metabolic plus redox homeostasis. This is also valid for other secretagogues, such as branched chain ketoacids (BCKAs); and partly for fatty acids (FAs). Branched chain aminoacids, leucine, valine and isoleucine, after being converted to BCKAs are metabolized by a series of reactions resembling β-oxidation of FAs. This increases superoxide formation in mitochondria, including its portion elevated due to the function of electron transfer flavoprotein ubiquinone oxidoreductase (ETF:QOR). After superoxide conversion to H2O2 the oxidation of BCKAs provides the mitochondrial redox signaling extending up to the plasma membrane to induce its depolarization together with the elevated ATP. In contrast, experimental FA-stimulated insulin secretion in the presence of non-stimulating glucose concentrations is predominantly mediated by GPR40, for which intramitochondrial redox signaling activates phospholipase iPLA2γ, cleaving free FAs from mitochondrial membranes, which diffuse to the plasma membrane and largely amplify the GPR40 response. These events are concomitant to the insulin release due to the metabolic component. Hypothetically, redox signaling may proceed by simple H2O2 diffusion or via an SH-relay enabled by peroxiredoxins to target proteins. However, these aspects have yet to be elucidated.

Keywords

  • pancreatic β-cells
  • insulin secretion
  • redox signaling
  • NADPH oxidase 4
  • branched chain ketoacid oxidation
  • fatty acid β-oxidation
  • ATP-sensitive K+ channel
  • GLP1
  • GPR40

1. Introduction

Recently, we revealed that physiological redox signaling is essential for the first phase of glucose-stimulated insulin secretion (GSIS) in pancreatic β-cells. Elevatedglucose intake contributes to the increasing pentose phosphate pathway (PPP) supply of NADPH for NADPH oxidase 4 (NOX4), which directly produces H2O2. The burst of H2O2 then represents a redox signal, which fundamentally determines GSIS, while inducing a cooperative induction of plasma membrane depolarization together with ATP elevation (Figure 1) [1]. The latter originates from the increased ATP synthesis by oxidative phosphorylation (OXPHOS). Hypothetically, either a closure of the ATP-sensitive K+ channel (KATP) is dependent on both H2O2 plus ATP; or H2O2 activates a synergic channel such as transient receptor potential melastin 2 (TRPM2) [2], required for sufficient depolarization. This principle of a logical summation of metabolic plus redox stimulation seems to be universal for other secretagogues (i.e. compounds stimulating insulin secretion), dependent on KATP. However, note that the redox signaling must be distinguished from the oxidative stress [3, 4, 5, 6].

Figure 1.

(A) Traditional (“standard”) view of the triggering mechanism of GSIS compared with (B) new paradigm in GSIS mechanism (“novel”), for which the redox signaling by NOX4-produced H2O2 is essentially required. Upon the glucose intake, PPP and redox shuttles supply cytosolic NADPH to increase NOX4 activity and thus elevate H2O2 which substantiates redox signaling. Either, the ATP-sensitive K+ channel (KATP) is closed exclusively when both ATP plus H2O2 are elevated. Alternatively, H2O2 activates opening of TRPM2 or other nonspecific cation channels required for a depolarization shift to reach a threshold potential of −50 mV, at which the voltage-sensitive Ca2+ channels (CaL) become open, thus starting to fire the action potential. Resulting Ca2+ influx into the cell cytosol allows a complex process of exocytosis of the insulin granule vesicles (IGVs), beginning during the so-called 1st phase of GSIS by fusion of pre-attached IGVs with the plasma membrane and exposure of the IGV interior to the extracellular space (capillaries in vivo). Ca2+ also promotes the recruitment of distant IGVs towards the plasma membrane as well as ensures the late, so-called 2nd phase of GSIS, lasting about 1 hour in vivo.

Advertisement

2. Two phases of GSIS

Two phases exist for GSIS in vivo [7, 8, 9, 10, 11]. They are also recognized in isolated pancreatic islets (PIs), but not in insulin-secreting β-cell lines. The consensus became that both KATP-dependent mechanism (also termed „triggering”) and KATP-independent mechanisms contribute to both phases [12], while the KATP-independent mechanism still requires the elevation of cytosolic Ca2+ [13]. The 2nd phase in vivo was even considered to be independent of the extracellular glucose concentrations [14]. It depends more on the molecular mechanism of the increased sustained mobilization and priming of insulin granule vesicles (IGVs) [15].

The first rapid peak of insulin secretion is observed at 5–10 min after administration of a bolus of glucose in vivo or addition of glucose to the isolated PIs. The 1st phase involves the exocytosis of pre-docked juxtaposed IGVs, residing 100–200 nm from the plasma membrane prior to triggering [16, 17] and also possesses a contribution of deeper localized granules arriving within 50 ms, which were not initially pre-docked [18, 19]. The 2nd phase typically lasts over 1 hr. As a result, a predominant insulin amount is released in this phase. The 2nd phase results most likely from further delayed recruitment of IGVs belonging to the typically excessive reserve. The past hypothesis suggested a main reason for such a delay to involve the restricted passage through the filamentous actin (F-actin) cytoskeleton [20, 21, 22], but later a microfilament-independent movement of IGVs was reported [21, 23, 24, 25]. However, numerous cytoskeleton components play a more detailed role in the IGV exocytosis, not representing only a simple barrier. Generally, the IGV exocytosis relies on synaptogamin activation by Ca2+, syntaxin, SNAP-25, and other target proteins of the SNARE family (SNAp REceptors, where SNAP is soluble NSF attachment proteins and NSF is a N-ethylmaleimide-sensitive fusion factor). They attract IGVs via the IGV-localized synaptobrevins (vesicle-associated membrane proteins), while forming a coiled-coil quarternaly structure [26]. The resulting SNARE core complex relocates the IGV and plasma membrane into proximity, thus facilitating establishment of so-called fusion stalk. Further zippering of coiled-coil structures allows fusion of larger part of the IGV membrane with the plasma membrane until a fusion pore is formed.

However, the recent explanation for the second phase is based on the fact that the two phases of insulin secretion exist when isolated pancreatic islets are studied, but do not exist for isolated primary pancreatic β-cells [27, 28, 29]. Hence, the role of inter-cellular contacts is emphasized for the 2nd phase. The inter-cellular contacts allow synchronization of the plasma membrane potential, while paracrine hormone secretion may also contribute to modification and termination of insulin release.

Advertisement

3. Mechanisms of the 1st phase of GSIS

GSIS has been consensually described to involve a so-called triggering mechanism accompanied by amplifying mechanism(s) [7, 12, 30, 31, 32, 33, 34, 35, 36]. The triggering is exclusively dependent on the KATP closure and attaining plasma membrane depolarization up to −50 mV. The latter is achieved in a synergy of KATP with other ion channels. The amplifying mechanisms are given by metabolism or stem from the action of incretins and other hormones. Also, mechanisms concerning other secretagogues, such as branched chain ketoacids (BCKAs) and fatty acids (FAs), were considered as merely amplifying. Nevertheless, we will show below the ambiguity of such a classification. The amplifying mechanisms originate from an incremental increase in Ca2+ elevations, not existing within the canonical “triggering” mechanism. Alternatively, they stem from facilitation via numerous proteins of the exocytotic machinery localized either on the IGV or plasma membranes. Therefore, some of these types of events might be Ca2+ independent and hence may also proceed at low glucose concentrations.

The traditional explanation of the triggering mechanism of GSIS relied exclusively on the ATP elevation (or elevation of the ATP/ADP ratio) in the cytosol of β-cells. Sole elevated ATP was considered to be sufficient for the KATP closing [30, 31, 32, 33]. Any additional requirement for a parallel redox signaling was not considered, despite the findings that reactive oxygen species (ROS) have been implicated in insulin secretion. This concerned with at least ROS of mitochondrial origin [37], or resulting from mono-oleoyl-glycerol addition [38]. The blockage of PPP, that decreased insulin secretion, also shifted redox homeostasis [39]. An unspecified link of GSIS with the externally added H2O2 was reported, besides antioxidant effects at decreased glutathione by diethylmaleate [40]. Previously, also an unidentified isoform of NADPH oxidase was implicated in GSIS, since an antisense p47PHOX oligonucleotide [41] or an unspecific NOX inhibitors attenuated GSIS [38, 42, 43].

Recently, we have provided the evidence that the elevated OXPHOS is insufficient to initiate GSIS, despite the increased ATP levels and the elevated ATP/ADP ratio at the peri-plasma- membrane space [1]. We demonstrated that NOX4 is fundamentally required for GSIS [1]. In model rat pancreatic β-cells (INS-1E cells) with silenced NOX4 or in full NOX4 knockout (NOX4KO) mice and in mice with NOX4 knockout, specifically in pancreatic β-cells (NOX4βKO mice), the 1st phase of GSIS was largely blocked [1]. In both studied NOX4 KO mice strains and in their isolated PIs, the 1st phase of GSIS was abolished with NOX4 ablation, while in PIs, either overexpression of NOX4 (achieved at least in the peripheral spheroid layer of islets) or additions of H2O2 rescued this 1st phase. No effects were found in NOX2 KO mice, although NOX2 has been previously implicated to play an antagonistic role for redox homeostasis [44].

Moreover, using a patch-clamp of INS-1E cells, we demonstrated that the KATP closure is possible only when NOX4 is intact in INS-1E cells. After showing the well-known closure of KATP induced by high glucose concentration in cells transfected with scrambled siRNA, we observed no glucose-induced KATP closure in INS-1E cells silenced for NOX4 [1]. These experiments supported the model, in which KATP integrates metabolic and redox homeostasis and acts as a logical summation for which both elevated ATP plus elevated H2O2 exclusively lead to a triggering of GSIS (Figure 1). However, since without cation fluxes provided by nonspecific cation channels a threshold depolarization of −50 mV cannot be achieved, despite 100% KATP ensemble being closed [45, 46], we may also hypothesize that H2O2 alternatively or in parallel activates the TRPM2 channel [2], known to contain redox-sensitive Met residue [47].

Thus our results set a new paradigm for GSIS, since it had never been considered that the sole ATP increase is insufficient for GSIS and is insufficient particularly for the closing of KATP; as well as it had never been considered that any redox signaling might essentially participate in GSIS.

In further work, we also demonstrated that the redox signaling upon GSIS is provided by elevations of cytosolic H2O2, whereas ROS in the mitochondrial matrix (both H2O2 and superoxide release) are diminished due to the enhanced operation of the redox shuttles upon GSIS [48]. One may expect that a portion of cytosolic NADPH as a substrate for NOX4 is provided by the glucose-6-phosphate dehydrogenase and also by 6-phosphogluconate dehydrogenase downstream within the PPP, whereas the second portion is generated due to the operation of redox shuttles. These shuttles become more active at higher glucose concentrations and increasingly produce NADPH. NADPH is particularly produced by isocitrate dehydrogenase 1 (IDH1) and malic enzyme 1 (ME1) in the cytosol upon operation of these redox shuttles [48].

In summary, we describe the revisited mechanism of the 1st phase of GSIS as follows. Elevated glucose metabolism and glycolysis allows an increased branching of the metabolic flux, particularly of glucose-6-phosphate G6P, toward PPP, which acts as a predominant source of NADPH. The essential role of PPP was emphasized elsewhere [49]. Amplification of the cytosolic NADPH is also provided by IDH1 and ME1 due to the elevated operation of the three redox shuttles. Since NOX4 was determined as the only NADPH oxidase producing H2O2 directly [50, 51], its reaction results in an increase of H2O2 release into the cell cytosol [1]. Finally, the elevated H2O2, together with concomitantly elevated ATP from the enhanced OXPHOS, is the only way for plasma membrane depolarization up to −50 mV [1]. This threshold subsequently induces CaL opening, followed by the Ca2+ influx into the cell cytosol, which in turn induces the exocytosis of insulin granule vesicles. The action potential spikes are then determined by the cycles of CaL opening, followed by the opening of voltage-dependent channels (KV) in rodents [52] or calcium-dependent (KCa) K+-channels in humans. Their action deactivates CaL, which are, however, again activated in the next CaL-KV cycle.

Pancreatic β-cells were undoubtedly adapted by phylogenesis to serve as a perfect glucose sensor. The glucose sensing is allowed by several key specific features. At first, specific isoforms of glucose transporters, GLUT2 in rodents and GLUT1 in humans, equilibrate the plasma glucose concentration with the glucose concentration in the cytosol of β-cells [53, 54]. Second, a specific isoform IV of hexokinase (also termed glucokinase) cannot be feed-back inhibited by its product glucose-6-phosphate. As a result, there is an efficient unidirectional flux towards the glycolysis [55, 56] and, most probably, this allows also branching into the PPP [49]. Originally, the PPP was accounted to utilize only 10% of glucose, due to presumably feed-back inhibition by glucose [57, 58]. However, metabolomics studies associated PPP intermediates with GSIS [59], confirming previous studies with various PPP inhibitors [59, 60, 61, 62]. These results collectively demonstrated the important PPP contribution to GSIS. This contribution is also reflected by existing patients having a deficiency of glucose-6-phosphate dehydrogenase associated with the impaired 1st phase of GSIS [63].

The third aspect leading to the perfect glucose sensing lays in the virtual absence of lactate dehydrogenase in β-cells and inefficiency in pyruvate dehydrogenase kinases (PDK) [64]. PDKs would otherwise block pyruvate dehydrogenase (PDH). Thus the highly active PDH and other dehydrogenases, activated also by Ca2+ influx into the mitochondrial matrix [65], altogether enable that 100% of pyruvate and its equivalents (after pyruvate conversion by transaminases) is utilized by OXPHOS. A minor pyruvate flux ensures anaplerosis of oxaloacetate due to the reaction of pyruvate carboxylase [66]. Its reaction is also important also for the pyruvate/malate redox shuttle.

The fourth aspect reflects the in vivo inhibitory role of the mitochondrial ATPase inhibitory factor, IF1. IF1 adjusts a proper glucose concentration range for GSIS in rat pancreatic β-cells, INS-1E [67, 68]. This is suggested by the demonstration that IF1 silencing allows insulin secretion even at very low glucose approaching to zero in INS-1E cells [67]. In contrast, the IF1 overexpression inhibited GSIS in INS-1E cells [68]. This IF1 role awaits confirmation in vivo.

Advertisement

4. Plasma membrane events following KATP closure

Surprisingly, the plasma membrane of β-cells contains up to 60 channels of 16 ion channel families [69]. Moreover, ion channels are also located on the membrane of IGVs to facilitate fusion with the plasma membrane and insulin exocytosis. Resting plasma membrane potential (Vp) is created predominantly by the activity of K+-channels due to a higher concentration of K+ inside the β-cell (~150 mM), exceeding that one established outside in capillaries or interstitial fluid (~5 mM). Experimentally, Vp values are measured to be approximately of – 75 mV [70]. The KATP closure then induces depolarization [69, 71, 72, 73] and activation of CaL [74]. The action potential firing is the entity that activates CaL-KV cycles (in rodents), however this firing is initiated by more channel types.

Surprisingly, the action potential firing is not induced until >90% of KATP channels are closed [75, 76]. As a result, only the closure of the remaining ~10% of the KATP population leads to depolarization [76]. In fact, the activity of the whole KATP population decreases exponentially with the increasing glucose concentration. Interestingly, 50% of the KATP population is already closed at 2–3 mM glucose, while Vp remains steady. However, at about 7 mM glucose, 100% of the KATP ensemble is closed. This is being reflected by the completely vanished KATP current, which leads to action potential firing [69, 70]. This event is termed as a supra-threshold depolarization.

Thus, hyperpolarized interburst phases are induced, while a nearly permanent firing exists at high >25 mM glucose [70]. An intermediate depolarization at 10 mM glucose was reported for mouse β-cells, reversed upon withdrawal of Ca2+ and Na+, supporting the participation of other channels, such as nonspecific cation channels, contributing to the depolarization (inward) flux [45]. Even an efflux of Cl was suggested to fulfil this role [77], including the opening of LRRC8/VRAC anion channels [78, 79]. The participation of TRPM4 and TRPM5 [80] providing inward currents of certain levels seem to be required for induction of sufficient membrane depolarization together with KATP closing [46]. This is because the measured resting Vp of – 75 to −70 mV is already depolarized by a some extent from the equilibrium Vpequi of −82 mV (5 mM vs. 130 mM [K+]). The shift is probably due to the opening of nonspecific cation channels, since any of Na+, Ca2+ and K+ can penetrate them. The 100% KATP closing at higher glucose causes only an insufficient depolarization. Without nonspecific cation channels (or Cl channels), the established Vp would only be equal to Vpequi, so any shift to −50 mV required for CaL would not take place. Contribution by the basal opening of other synergic channels is therefore essential. Open synergic channels always induce the inward shift in Vp, so to that depolarization given by 100% KATP closing reaches −50 mV. This allows opening of CaL and action potential firing. In summary, besides the heat-activated TRPV1 channel (capsaicin receptor), and TRPV2 or TRPV4, the H2O2-activated TRPM2 [2], or Ca2+-activated TRPM4 and TRPM5 channels belong to the important group of possible synergic channels expressed in β-cells [46].

The same reasoning concerns with anion channels, particularly Cl channels. The active Cl transport is provided in β-cells by SLC12A, SLC4A, and SlC26A channels. These channels set the cytosolic Cl concentration above thermodynamic equilibrium. Besides GABAA, GABAB and glycine receptor Cl channels considered to be part of the insulin secretion machinery, also volume-regulated anion channels (VRAC) were shown to be open at high glucose. VRACs are heteromers of the leucine-rich repeat containing 8 isoform A (LRRC8A) with other LLRC8 isoforms, forming anion channels [79]. Ablation of LRRC8 in mice led to delayed Ca2+ responses of β-cells to glucose and diminished GSIS in mice, demonstrating the modulatory role of LRRC8A/VRAC on membrane depolarization leading to CaL responses [78, 79].

Upon the action potential firing thus metabolically driven Vp oscillations occur due to the initial glucose rising [69, 70]. Cytosolic Ca2+ oscillations are superimposed from fast (2–60 s periods) and slow (up to several min) Ca2+ oscillations [81], stemming from Vp oscillations and an interplay with Ca2+ efflux from the endoplasmic reticulum (ER) [82]. Collectively they lead to pulsatile insulin secretion. The ER involvement is given by the phospholipase C (PLC), responding to the glucose-stimulated Ca2+ influx. PLC produces inositol triphosphate (IP3), which opens the Ca2+ channel of IP3 receptor (IP3R) of ER; plus diacylglycerol (DAG). Importantly, DAG permits the opening of TRPM4 and TRPM5 via the protein kinase C (PKC) pathway. Another ER Ca2+ channel, the ryanodine receptor (RyR) may also participate, being activated by ATP, fructose, long-chain acyl-CoAs and cyclic adenosine 5′-diphosphate ribose [81]. Also, the role of other channels was demonstrated for permitting store-operated Ca2+ entry from ER, particularly of the ternary complex of TRPC1/Orai1/STIM1 [46, 83]. TRPC1 belongs to the transient receptor potential canonical (TRPC) family with a modest Ca2+ selectivity. TRPC1 interacts with Orai1 [84], and in such a functional complex, its channels are activated by STIM1, affecting the amplitude of Ca2+ oscillations, and correlating with GSIS.

As mentioned above, deactivation of CaL is ensured by the opening of voltage-dependent channels (KV) in rodents [52] or calcium-dependent (KCa) K+-channels in humans. Among the former, tetrameric KV2.1 is the prevalent form in rodent β-cells. A delayed rectifier K+-current is induced at positive Vp down to −30 mV [85]. The opening of KV2.1 channels repolarizes Vp and thus closes CaL channels. Ablation of KV2.1 thus reduces Kv currents by ~80% and prolongs the duration of the action potential, so more insulin is secreted. Mice with ablated KV2.1 possess lower fasting glycemia but elevated insulin and reportedly improved GSIS [86]. In contrast, human β-cells use KCa1.1 channels (i.e. BK channels) for repolarization of Vp [70]. Note also that downregulation of KV was observed after islet incubation with high glucose for 24 hr [87].

Advertisement

5. Possible redox regulations of KATP and other channels

The structure of KATP has been resolved and numerous mutagenesis studies of KATP have been conducted. Amino acid residues that are candidate redox targets are yet to be identified. The KATP channel is a hetero-octamer consisting of four external regulatory sulfonylurea receptor 1 (SUR1, a product of Abcc8 gene) subunits and four pore-forming subunits of potassium inward rectifier, Kir6.2 (Kcnj11 gene) [88, 89]. These Kir6.2 subunits cluster in the middle of ~18 nm size structure with a ~13 nm height [90]. The part exposed to the cytosol contains an ATP binding site, located about 2 nm below the membrane. A single ATP molecule was reported to close the channel, i.e. with the other three binding sites left unoccupied [91]. However, the ATP binding site overlaps with the binding site for phosphatidylinositol 4,5-bisphosphate (PIP2), which stabilizes the open state. Palmitoylation of Cys166 of Kir6.2 was then reported to amplify the responsiveness to PIP2 [92]. Upon the release of PIP2 from the binding site, the open probability becomes decreased [90, 93, 94].

Diazoxide or cromakalim, as well as numerous other openers, set KATP pharmacologically in the open state even at a high ATP concentration [95]. In contrast, the artificial KATP closing by sulfonylurea derivatives, such as glibenclamide, takes place independently of ATP. Besides this sulfonylurea binding site, each of the four SUR1 subunits contains MgATP and MgADP binding sites. MgATP is hydrolyzed at the nucleotide binding fold 1 (NBF1) to MgADP. Resulting MgADP subsequently activates KATP at NBF2. This is indeed reflected by the ATP-sensitive increase in K+ conductance and following lower excitability, accompanied by the lower sensitivity to ATP inhibition [91].

The phosphorylation of KATP reportedly sets the sensitivity of the KATP ensemble. The setting is such that transitions upon the glucose rise from 3 or 5 mM to 7 mM or > 10 mM result in the closing of the remaining 10% of the initially open channels by elevations between just the two ATP concentrations falling into the mM range. Any redox component in this was never indicated and should be studied. Nevertheless, phosphorylation mediated by the protein kinase A (PKA) was already reported to act in this unusual setting. Thus Thr224 [96] and Ser372 were reported to be the verified PKA phosphorylation sites. Their phosphorylation increases the open probability of KATP [97]. This might hypothetically provide closing mechanism acting at higher ATP concentration or even requiring H2O2. In a longer time scale, phosphorylation also increases the number of channels in the plasma membrane. Also, Thr224 was found to be phosphorylated by Ca2+ and calmodulin-dependent kinase II (CaMKII) while interacting with βIV-spectrin [98]. In vivo, also autonomic innervations and paracrine stimulation ensure sufficient PKA-mediated phosphorylation of KATP.

Since the original discovery of the essential role of KATP in GSIS [99], only an indirect inhibition of KATP by H2O2 was observed in smooth muscle cells [100]. Nevertheless, other redox-sensitive targets have been identified in pancreatic β-cells. But we can exclude the possibility that the IGV exocytosis itself might be directly induced by H2O2, independently of Ca2+ [52], since the ability of exogenous H2O2 to induce insulin secretion in INS-1E cells was only partially blocked by NOX4-siRNA, but it was completely blocked by a CaL blocker nimodipine [1]. Consequently, albeit the used H2O2 doses exceeded 100 μM, they did not directly stimulate the KATP-independent exocytosis of insulin granules.

A second possibility would be that CaL channels themselves may be hypothetically co-activated by H2O2. Third, the plasma membrane depolarization might be redox sensitive, so that H2O2 could directly or indirectly inhibit repolarizing K+-channels, such as KV [101, 102, 103]. The fourth plausible redox link with GSIS would concern with the reported redox activation of TPRM2 depolarizing channels [2]. The latter is the most plausible, since it is related to a Ca2+-induced [52, 104] or H2O2-induced exocytosis of insulin granules by the H2O2-activation of TPRM2 depolarizing channels [2, 105]. Note, our results excluded the Ca2+-independent H2O2-induced exocytosis of IGVs at least in rat pancreatic β-cells [1]. Therefore, if the H2O2-activated TRPM2-dependent mechanism exists, it must provide the required synergy with KATP, to reach the −50 mV plasma membrane depolarization threshold. Note also, that TRPM2 was already implicated as a significant player in the GLP-1 potentiation of insulin secretion [106].

Finally, a competition for NADPH between NOX4 and a hypothetical NADPH-activated K+-channel could exist. Nevertheless, using patch-clamped INS-1E cells in a whole cell mode, we demonstrated a closure of KATP by H2O2 produced by NOX4 at high glucose, since in cells silenced for NOX4, even ATP resulting from the metabolism of high glucose was unable to close the KATP channel [1].

Advertisement

6. Receptor-mediated amplification of insulin secretion

G protein-coupled receptors activating heterotrimeric G proteins ensure pleiades of cell responses, mutually interrelated. G proteins typically regulate production of second messengers. Thus Gαs proteins increase generation of cyclic AMP (cAMP), whereas Gαi/o proteins decrease it [107, 108, 109]. The G proteins Gαq/11 initiate PLC-mediated hydrolysis of phosphatidylinositol 4,5-bisphosphate into diacylglycerol (DAG) and IP3 [110, 111]. Gα12/13 proteins promote protein RhoA for remodeling of the cytoskeleton [112]. Class of proteins termed β-arrestins initiates signaling via proximal MAP kinase, IκB, and Akt pathways [113]. The latter two G protein classes rather control long-term effects.

Let us emphasize downstream pathways that are important for acute effects in pancreatic β-cells, which predominantly lead to either modulation of the plasma membrane channels, typically CaL, KATP and KV, so to ensure more intensive insulin secretion; or their action evokes stimulation of insulin secretion via ensuring the surplus Ca2+ influx to the cytosol from ER or mitochondria; or, else, their action targets proteins of the exocytotic machinery on the IGV or plasma membranes. The latter responses alter the kinetics of IGVs in docking, priming and fusion with the plasma membrane, so to facilitate exocytosis. Interestingly, these events could be independent of CaL and theoretically could take place at low glucose concentrations.

Activation of Gαs increases the activity of transmembrane adenylate cyclases (tmAC) producing cAMP from ATP [108, 109]. A number of phosphodiesterases (of 11 families) degrade cAMP (some also or exclusively cGMP). cAMP is a universal 2nd messenger having a specific function in amplifying of GSIS and insulin secretion stimulated with other secretagogues. Also, soluble adenylate cyclases (sACs) exist, notably in the mitochondrial matrix, while their reaction is potentiated by Ca2+ and bicarbonate. The major mediators of cAMP effects are cAMP-dependent PKA [114], including PKA tethered to the outer mitochondrial membrane [115, 116], and the parallel pathway of enhanced signaling via exchange proteins directly activated by cAMP 2 (EPAC2) [117, 118, 119].

In pancreatic β-cells, the PKA pathway is involved in signaling of incretin (GLP-1 and GIP) receptors [107, 120]. It exerts a minor contribution to signaling from metabotropic receptors, such as GPR40, which is sensing long chain fatty acids [111]. PKA typically amplifies the Ca2+-dependent exocytosis of insulin granules. The core pathway involves PKA phosphorylation and hence activation of the CaL β2-subunit, in concert with KATP phosphorylation decreasing the ATP concentration range required for its closure (see above) [121]. In addition, PKA inhibits Kv channels, which otherwise terminate plasma membrane depolarization; hence this prolongs already more intensive Ca2+ influx via phosphorylated CaL and hence exocytosis of insulin granules [122].

Another PKA target is the exocytosis-modulating protein termed snapin, the phosphorylation of which allows its interaction with the other IGV proteins, which enhances the 1st GSIS phase [123]. Snapin participates in tethering of IGVs to the plasma membrane by coiled-coil interaction with a lipid-anchored protein SNAP-25 [124].

Altogether, the PKA pathway ensures about 50% of cAMP responses in β-cells [125], while the EPAC2 pathway ensures the remaining responses [117, 118, 119]. EPAC2 protein possesses a guanine nucleotide exchange activity, thus inducing the Ca2+-induced Ca2+release from ER via RyR [126] (questioned in [127]), occurring only at high glucose, since it requires the primary CaL opening [128], which also partially refills the ER Ca2+ stores. The EPAC2 pathway also affects the IGV proteins and thus facilitates the insulin exocytosis. For example, Rim2a protein is a target [129130], located on the inner plasma membrane surface and on IGVs, representing a scaffold for IGV exocytosis [131]. Rim2a interacts with Rab3A of IGVs and the resulting Rim2a-Rab3A complex facilitates docking of IGVs into the plasma membrane. This is followed by so-called priming, which is subsequently initiated by the Rim2a interaction with the Munc13–1 protein. Munc13–1 then opens syntaxin 1 from its closed conformation, thus allowing fusion with the plasma membrane. EPAC2 also interacts with NBD1 of SUR1, being released by cAMP [35]. Such locally released EPAC2 induces the release of Rim2 from the α1.2 CaL subunit. The local Ca2+ influx within CaL ensures EPAC2 binding to Rim2, and subsequent interaction with another Ca2+ sensor termed Piccolo. The heterotrimeric complex then interacts with Rab3A and enables IGV exocytosis.

Interestingly, all necessary components of the PKA pathway were identified in the mitochondrial matrix, including sAC, PDE2A2 [132], and also PKA [133]. However, we may also speculate that some proteins can be phosphorylated by cytosolic PKA or by its fraction attached to OMM prior to their import to the mitochondrial matrix. There was also a consensus that cAMP cannot freely diffuse to the matrix [132]. Thus cAMP in the mitochondrial matrix may act as an independent pool [134, 135]. Its source is the matrix-located soluble adenylate cyclase sAC, which is activated by bicarbonate and Ca2+ [136, 137]. Since CO2 is increasingly released when the Krebs cycle turnover increases upon GSIS, the matrix localized mtPKA can be activated in this way [138]. In any case, OXPHOS is facilitated in mitochondria of numerous tissues via phosphorylation of Complex I NDUFS4 subunit (facilitating its Hsp70-mediated import), Complex IV COXIV-1 subunit (preventing its inhibition by ATP) [139] as well as via IF1, enhancing ATP synthesis by disabling the inhibitory binding of phosphorylated IF1 dimers to the ATP synthase [140]. A link to redox homeostasis can be viewed in the observed release of the PKA catalytic subunits by the increased ROS [141, 142]. Thus mtPKA can act in parallel to the cytosolic PKA signaling initiated by GPR40 and GLPR or GIPR receptors. PKA targeting of at least IF1, and probably also of Complex I and Complex IV, should contribute to the amplification of insulin secretion by FAs or incretins.

The G protein Gαq/11 initiates signaling through the phospholipase C (PLC-)-mediated hydrolysis of phosphatidylinositol 4,5-bisphosphate into DAG and inositol triphosphate IP3 [110]. The main effector of DAG is protein kinase C (PKC), which is activated by DAG. One of the effectors of IP3 is the IP3 receptor (IP3R; subtypes IP3R1, IP3R2 and IP3R3), which is another important Ca2+ channel residing on ER membranes in β-cells [143]. Similarly to the EPAC2-RyR route of Ca2+ release from ER Ca2+, the opening of this channel amplifies the primary CaL mediated Ca2+ signaling for insulin release. PKC contributes to the plasma membrane depolarization, while activating TRPM4 and TRPM5 [144]. Besides the canonical plasma membrane effects, PKC and downstream ERK1/2 signaling stimulates OXPHOS, hence mitochondrial ATP synthesis [145].

Advertisement

7. GSIS amplification by incretins GLP-1 and GIP

Glucagon-like peptide 1 (GLP-1) and gastric inhibitory polypeptide (GIP) have a prominent impact among other peptides belonging to incretins [107, 108, 109]. Oral glucose administration provides a higher insulin secretion response than when administered parenterally [146]. This surplus of potentiation of insulin secretion appears to be about equally ascribed to GLP-1 and GIP [147]. Indeed, diminished insulin secretion response to oral glucose was observed in GLP-1 knock out mice [148, 149] and was even more decreased in double knockout mice (GLP-1 plus GIP) [149].

Incretin-cAMP signaling amplifies GSIS by both PKA-dependent and EPAC2A-dependent pathways. As described above, the EPAC2 pathway is partially dependent on the CaL opening, and the PKA pathway enables synergy among actions of KATP, CaL, and Kv channels, leading again to a more effective CaL opening. This knowledge complies with the traditional view, considering that the incretin signaling does not stimulate insulin release in the low glucose conditions [150, 151]. The GLP1R-cAMP-EPAC2-TRPM2 pathway was suggested to be one of the major routes [106].

GLP-1 is secreted by enteroendocrine L-cells, residing predominantly in the distal ileum and colon. Secretion is initiated by postprandial stimuli, i.e. by glucose, fatty acids, or lipids, as well as proteins [152, 153]. Only 10 to 15% of active GLP-1 likely reaches the pancreas via the circulation [154]. Thus concentrations of biologically active GLP-1 in human plasma at fasting account for about 2 pmol/l and maximum10 pmol/l postprandially [155], peaking 30 to 60 min after a carbohydrate or protein intake and 120 min after ingestion of lipids [156]. The most efficient truncated variants are GLP-1(7–37) and variant GLP-1(7–36amide) [152]. The latter is ~80% abundant in humans [157]. Note that full peptide GLP-1(1–37) is much less efficient in GSIS potentiation [150, 151]. Moreover, paracrine GLP-1 signaling acts among the different types of PI cells [150], similarly to the paracrine and endocrine secretion of other hormones. On the systemic level, central control by the brain and nervous system, including GLP-1 secretion in the nucleus tractus solitarii of the brainstem [152], further provides an indispensable top level of regulation for the insulin secretion. GLP-1 effects related to β-cell proliferation or apoptosis are beyond the scope of this review.

GLP-1 from the bloodstream acts through its receptor (GLP1R) residing in the plasma membrane of pancreatic β-cells [158]. GLP1R activation stimulates Gαs and Gαq/11 and recruits β-arrestin, depending on biased agonism relative to different agonists, such as exendin-4 and oxyntomodulin [159, 160]. As a scaffold protein, β-arrestin facilitates signaling via Gαs to cAMP but also to CREB [160], extracellular regulated kinase ERK1/2 [161], and insulin receptor substrate 2 (IRS-2), the effects promoting β-cell growth, differentiation, and maintenance [160]. The stimulation of Gαs leads via enhanced cAMP to the initiation of PKA [162] and EPAC2A pathways [163]. Continuous cAMP production and partial potentiation of GSIS was found even for the internalized GLP1R [164].

The PKA pathway provides a surplus intracellular Ca2+ above that of the net GSIS without any receptor stimulation. This is ensured by phosphorylation-induced closing among the population of KATP, stimulation of CaL opening, and closing of Kv channels [165]. The latter prolongs Ca2+ stimulation of IGV exocytosis and hence may also potentiate the 2nd phase of GSIS. In parallel, PKA engages snapin interaction with IGVs, reportedly potentiating the 1st GSIS phase [123, 124]. Simultaneously, the EPAC2 pathway promotes Ca2+-induced RyR-mediated Ca2+ release from ER, which must be, however, initiated by the ongoing CaL opening [163]. The EPAC2 pathway also facilitates docking and priming of IGVs by promoting Rab3A interaction with Rim2a [131] and hypothetically interaction of EPAC2-Rim2-Picollo trimers with Rab3A, enabling IGV exocytosis [152]. Stimulation of GLP1R biased downstream via stimulation of Gαq/11 also contributes by a surplus to intracellular Ca2+, while inducing the IP3R-mediated Ca2+ release from ER.

When GLP1 effects were simulated and IGV kinetics was monitored using total internal reflection fluorescence microscopy, cAMP and 8-Br-cAMP were found to increase the frequency of fusion events, i.e. IGV fusion with the plasma membrane in both phases of GSIS [25]. EPAC2A was found to interacts also with a small G protein Rap1, affecting its conformation so to release the catalytic region, which subsequently binds and thus activates another G protein Rap113. In EPAC2A knockout mice, most of the potentiation of the 1st GSIS phase vanished [25]. Thus speculatively, the 2nd phase amplification can be due to the PKA pathway.

Advertisement

8. Mechanism of insulin secretion stimulated by branched-chain keto-acids

Postprandial response by insulin secretion is also given by substances other than glucose. These substances, which induce the secretion of insulin, are termed secretagogues in general. One important type of secretagogues is branched-chain keto-acids (BCKAs), metabolites of branched-chain amino acids (BCAAs) (Figure 2). We found that the alternative to the NOX4-mediated redox signaling exists for some other insulin secretagogues, particularly for BCKAs [1]. For the redox signaling in this case, the mitochondrial redox signaling replaced that one originating from NOX4. Thus we demonstrated that H2O2 signaling originating from mitochondria is essentially required for insulin secretion stimulated by BCAAs metabolized onto BCKAs, such as 2-ketoisocaproate (KIC; also termed 2-oxoisocaproate, OIC; leucine metabolite), 2-ketoisovalerate (KIV; valine metabolite) and 2-ketomethylvalerate (KMV; isoleucine metabolite) [166, 167]. This mechanism was evidenced by the effects of mitochondrial-matrix-targeted antioxidant SkQ1. We observed that SkQ1 did not affect GSIS in INS-1E cells, but completely inhibited insulin secretion stimulated by KIC [1]. Thus the NOX4 source of H2O2 cannot be efficiently inhibited by SkQ1 located within the inner phospholipid leaflet of the inner mitochondrial membrane, whereas the redox signaling originating from mitochondrion must be blocked.

Figure 2.

The mechanism of branched chain keto acid-stimulated insulin secretion involves redox signaling of mitochondrial origin. 2-ketoisocaproate (KIC), 2-ketomethylvalerate (KMV) and 2-ketoisovalerate (KIV) resulting from leucine, isoleucine, and valine, respectively, due to the branched chain aminotransferase reaction in mitochondria (BCAT2), are metabolized by the branched chain ketoacid dehydrogenase (BCKDH) in mitochondria. A series of reactions, BCKA oxidation leads to the electron transfer from the co-factor FADH2 of BCKDH, via ETF towards the ETF:QOR reaction reducing Q to QH2. This effectively retards the competing reaction of the Complex I of the mitochondrial respiratory chain, leading to the superoxide formation. In the mitochondrial matrix, superoxide is transformed to H2O2 by MnSOD, whereas by CuZnSOD in the intermembrane space and cytosol. The elevated mitochondrial/cytosolic H2O2 substitutes the redox signal of NOX4 origin. Consequently, such redox signaling, together with elevated ATP, allows the sufficient depolarization of the plasma membrane.

Metabolism of BCKAs begins in the mitochondrial matrix by the reaction of the BCKA dehydrogenase complex (BCKDH), since there is no branched-chain amino acid aminotransferase (BCAT) in the cell cytosol [168]. BCKDH forms isovaleryl-CoA, isobutyryl-Co and methyl-isobutyryl-CoA from KIC, KIV and KMV, respectively. This is followed by a series of reactions resembling β-oxidation of fatty acids. This series, as well as FA β-oxidation, elevates formation of superoxide in the mitochondrial matrix by several ways. The major way is due to the reoxidation of the BCKDH co-factor FADH2 by the electron-transfer flavoprotein (ETF), the one electron carrier. Two electrons from the two ETF molecules are accepted by the electron-transfer flavoprotein: ubiquinone oxidoreductase (ETF:QOR) [169]. ETF:QOR reaction is coupled to ubiquinone (Q) oxidation to ubiquinol (QH2). This effectively competes with the Complex I reaction of the respiratory chain, also providing Q oxidation to QH2 driven by NADH. The electron transfer within the Complex I is thus effectively retarded and this results in a higher superoxide formation. Superoxide is then most probably increasingly formed at the IQ site (i.e. at the proximity of the Q-binding site), similarly as due to the reverse electron transfer.

Alternatively, the Complex I electron transfer is retarded upon the acetyl-CoA entry (propionyl-CoA entry for KIV; through methylmalonyl and Succinyl-CoA) into the Krebs cycle. Also, acetoacetate influences redox homeostasis, as one of the final products of leucine metabolism. After superoxide conversion to the elevated H2O2 in the mitochondrial matrix, the H2O2 is elevated in the cytosol and thus represents the mitochondrial retrograde redox signaling. Its target could be again KATP (or KATP and TRPM2) which would depolarize the plasma membrane due to this redox signaling ongoing in parallel with the elevated ATP due to concomitantly enhanced OXPHOS.

The BCAA oxidation involves the following sequence of reactions: isovaleryl-CoA dehydrogenase (IVD), methylcrotonyl-CoA carboxylase (MCC), methyl-glutoconyl-CoA hydratase (MGCoAH) and 3-hydroxy-3-methylglutaryl-CoA lyase (HMGCoAL). The end-products are acetyl-CoA and acetoacetate. Similarly, as for pyruvate metabolism via PDH, the common end-product acetyl-CoA drives the Krebs cycle. This may also increase mitochondrial superoxide formation. Acetyl-CoA is linked to the above acyl-CoA dehydrogenase reaction by the reaction of the ETF:QOR, using ubiquinone (CoQ or Q) to oxidize it to ubiquinol QH. Also, the ETF:QOR itself may produce superoxide.

In summary, independently of the molecular mechanism, BCAA metabolism leads to the increased mitochondrial superoxide formation. After conversion to H2O2 by the matrix MnSOD and the intermembrane space CuZnSOD, the ongoing H2O2 efflux from mitochondria can be regarded as redox signaling. We have clearly demonstrated that the absence of such redox signaling, for example, in the presence of the mitochondrial matrix-targeted antioxidants SkQ1 leads to a blockage of insulin secretion, which is otherwise stimulated with BCKAs [1]. Likewise, the silencing of BCKDH led to the inhibition of insulin secretion stimulated with BCKAs.

Advertisement

9. Mechanism of fatty acid-stimulated insulin secretion

Fatty acids (FAs) appear in pancreatic islet capillaries either bound to albumin or being part of postprandial chylomicrons resulting from dietary fat lipids. FA pool of lipoproteins can be also considered. The dietary fat lipids are rich in triglycerides, which are cleaved locally in pancreatic islet capillaries by lipoprotein lipase secreted by β-cells. Resulting 2-monoacylglycerol (2MAG) and long chain FAs [170, 171, 172, 173] stimulate each own two receptors GPR119 [157] and GPR40/FFA1, respectively. Therefore, fatty acid-stimulated insulin secretion (FASIS) could be defined as the net insulin secretion induced at the low glucose concentration, which itself does not stimulate insulin secretion. It is still controversial, whether such a net FASIS exists, since some previous reports observed that glucose should always be present for fatty acid to induce insulin secretion response. In contrast, the other reports described FASIS at 3 mM glucose, but not at zero glucose. Physiologically, postprandial responses should be due to all secretagogues resulting from major saccharide, fat and protein components. FASIS may dominate late responses upon feeding by fatty meal or an experimental high-fat diet [174].

Theoretically, FASIS must concern with the two components (Figure 3). The first one should depend on metabolism and the second one should rely on the stimulation of the metabotropic receptor GPR40. In vivo, FASIS-GPR40 axis is paralleled by a portion of insulin secretion stimulated via another metabotropic receptor, GPR119, to which monoacylglycerol (MAG) binds as the second major component of triglycerides. The metabolic component undoubtedly involves fatty acid β-oxidation, providing both ATP from the elevated OXPHOS and H2O2 from the enhanced superoxide formation by the respiratory chain and ETF:QOR, similarly as for BCKAs [175]. This component is thus directly dependent on KATP, since it leads to its closure and to the canonical downstream events identical to those during GSIS.

Figure 3.

(A) Traditional (“standard”) view of FASIS compared with (B) FASIS with GPR40 receptor supplied by the mitochondrial fatty acids mechanism (“novel view”). The mechanism of FASIS at low glucose appears to be predominantly mediated by GPR40. Its ligands are free long-chain FAs. An excessive supply of GPR40 ligands and hence a substantial amplification of the GPR40 downstream response is given by the redox-activated mitochondrial phospholipase iPLA2γ/PNPLA8. The phospholipase iPLA2γ is directly activated by the elevated H2O2 in the mitochondrial matrix and cleaves both saturated and unsaturated FAs from the phospholipids of mitochondrial membranes. The cleaved free FAs diffuse up to the plasma membrane, where they activate GPR40.

The receptor component of FASIS may be at least partly KATP-independent and even CaL-independent, hence may partly proceed independently of high glucose. In other words, FASIS at low glucose is theoretically possible. The major pathway downstream of GPR40 relies on Gαq/11, which induces PLC-mediated hydrolysis of phosphatidylinositol 4,5-bisphosphate into DAG and IP3 [110, 111]. The latter would amplify the primary CaL-mediated Ca2+ signaling for the insulin release by mediating Ca2+ release from ER via the Ca2+ channel of the IP3 receptor [143]. However, this would happen provided that some basal CaL would be initiated by the metabolic component, i.e. due to partial fatty acid metabolism by β-oxidation followed by H2O2 plus ATP elevations. Also, PKC could be activated downstream of GPR40 as being the main effector of DAG. The PKC pathway could increase the extent of plasma membrane depolarization since it activates TRPM4 and TRPM5 [144]. Moreover, this may act at low glucose, again providing that the initial triggering is ensured by the metabolic component, and so that certain basal H2O2 plus ATP elevations exist, leading to the KATP closure. Also, another route downstream of GPR40 would involve the Gαq/11-PLC-TRPC-induced Ca2+ efflux from ER [176]. As mentioned above, the TRPC1-Orai1-STIM1 complex was demonstrated to act during GSIS, while contributing to Ca2+ oscillations [46, 83, 84]. The action of such a complex has yet to be studied during experimental FASIS as well as its dependence on CaL.

Also, biased (promiscuous) pathways of GPR40, i.e., those involving GαS-cAMP initiation of information signaling may exist and contribute to a certain extent to FASIS. Both downstream pathways of GPR40-GαS-cAMP stimulation, i.e. the PKA and EPAC2 pathway, could target components of IGV interactions with the plasma membrane, hence being independent of CaL. These speculations await experimental evidence.

Our in vitro and in vivo experiments with mice (unpublished) demonstrated that approximately 2/3 of the GPR40 response (amplitude of insulin secretion) is given by the amplifying mechanism due to the mitochondrial phospholipase iPLA2γ/PNPLA8 [175]. This phospholipase cleaves both saturated and unsaturated FAs from the phospholipids of mitochondrial membranes. The cleaved free FAs subsequently diffuse up to the plasma membrane, where they activate GPR40. Moreover, the phospholipase iPLA2γ is directly activated by the elevated H2O2 in the mitochondrial matrix. The reader may remain that this is just the FA β-oxidation, which via the increased superoxide formation, due to the function of ETF:QOR and respiratory chain, produces H2O2, while the concomitant OXPHOS provides elevated ATP. As a result, the sufficient plasma membrane depolarization is enabled. The proof of co-existence of the GPR40 receptor component and metabolic component of FASIS is suggested by the experiments when FASIS in iPLA2γ knockout mice or in its isolated islets yielded only ~30% insulin secretion peak in the 1st phase when compared to wt mice (Holendová B., Jabůrek M, et al., unpublished). Incidentally, a similar portion remains when GW1100 antagonist of GPR40 was applied. These results show that in parallel with the GPR40 pathway, a 1/3 portion of FASIS still results from FA β-oxidation, having a similar mechanism as described for ketoacids. The abolished FASIS in the iPLA2γ knockout mice then supports the existence of such an acute mechanism in vivo, when GPR40 is supplied with mitochondrial fatty acids.

Advertisement

10. Redox relay as a hypothetical carrier for redox signaling

It has been established that the content of glutathione (GSH) is rather low in pancreatic β-cells [177, 178, 179, 180], in contrast to the content of thioredoxins and glutaredoxins [181, 182], peroxiredoxins and other proteins capable of redox relay. Therefore, these proteins are able to conduct and spread the redox signals [183184]. From this point of view, the pancreatic β-cell appears to be a well-integrated redox system.

Redox signal spreading may be accomplished either by the direct diffusion of H2O2 or may be facilitated by the specialized proteins. Redox signals can be traced experimentally as instantly oxidatively modified cysteine residues, which are spread via different sets of proteins in different tissues. However, one may consider their majority as passive targets. For the case of NOX4 residing in the proximity of KATP, undoubtedly, the direct diffusion of H2O2 would be sufficient. Nevertheless for more distant NOX4 molecules, this would be difficult. Also, for mitochondrial redox signaling towards targets residing in the plasma membrane, a distance over 500 nm must be overcome. Redox signal across such high distances could be conducted through the action of thiol-based proteins capable of redox relay to the target, such as peroxiredoxins (regenerated via thioredoxins and glutaredoxins). The relay would provide a common redox signal transfer. It is yet to be established whether a redox relay exists via an array of peroxiredoxin oligomers.

Acknowledgments

This chapter is supported by grant No. 20-00408S from the Grant Agency of the Czech Republic (Grantová Agentura České Republiky).

References

  1. 1. Plecita-Hlavata L, Jaburek M, Holendova B, Tauber J, Pavluch V, Berkova Z, et al. Glucose-Stimulated Insulin Secretion Fundamentally Requires H2O2 Signaling by NADPH Oxidase 4. Diabetes. 2020. DOI: 10.2337/db19-1130
  2. 2. Grupe M, Myers G, Penner R, Fleig A. Activation of store-operated I(CRAC) by hydrogen peroxide. Cell Calcium. 2010;48:1-9. DOI: 10.1016/j.ceca.2010.05.005
  3. 3. Stoian AP, Mitrofan G, Colceag F, Suceveanu AI, Hainarosie R, Pituru S, et al. Oxidative stress in diabetes. A model of complex thinking applied in medicine. Revista de Chimie. 2018;69:2515-2519. DOI: 10.37358/RC.18.9.6566
  4. 4. Picu A, Petcu L, Ştefan DS, Grădișteanu Pîrcălăbioru G, Mitu M, Bajko D, et al. Evolution of Inflammatory and Oxidative Stress Markers in Romanian Obese Male Patients with Type 2 Diabetes Mellitus after Laparoscopic Sleeve Gastrectomy: One Year Follow-Up. Metabolites. 2020;10. DOI: 10.3390/metabo10080308
  5. 5. Pantea Stoian, A.M., G.; Colceag, F.; Serafinceanu, C.; Eftimie Totu, E.; Mocanu, V.; Mănuc, D.; Mihaela Cărăuşu, E. Oxidative Stress Applied in Diabetes Mellitus-A New Paradigm. Proceedings2019, 11, doi:10.3390/proceedings2019011007
  6. 6. Zhang P, Li T, Wu X, Nice EC, Huang C, Zhang Y. Oxidative stress and diabetes: antioxidative strategies. Frontiers of medicine. 2020. DOI: 10.1007/s11684-019-0729-1
  7. 7. Straub SG, Sharp GW. Glucose-stimulated signaling pathways in biphasic insulin secretion. Diabetes/Metabolism Research and Reviews. 2002;18:451-463. DOI: 10.1002/dmrr.329
  8. 8. Hoang DT, Hara M, Jo J. Design Principles of Pancreatic Islets: Glucose-Dependent Coordination of Hormone Pulses. PLoS One. 2016;11:e0152446. DOI: 10.1371/journal.pone.0152446
  9. 9. Kalwat MA, Cobb MH. Mechanisms of the amplifying pathway of insulin secretion in the β cell. Pharmacology & Therapeutics. 2017;179:17-30. DOI: 10.1016/j.pharmthera.2017.05.003
  10. 10. Villard O, Brun JF, Bories L, Molinari N, Benhamou PY, Berney T, et al. The Second Phase of Insulin Secretion in Nondiabetic Islet-Grafted Recipients Is Altered and Can Predict Graft Outcome. The Journal of Clinical Endocrinology and Metabolism. 2018;103:1310-1319. DOI: 10.1210/jc.2017-01342
  11. 11. Henquin JC, Dufrane D, Kerr-Conte J, Nenquin M. Dynamics of glucose-induced insulin secretion in normal human islets. American Journal of Physiology. Endocrinology and Metabolism. 2015;309:E640-E650. DOI: 10.1152/ajpendo.00251.2015
  12. 12. Henquin JC. Regulation of insulin secretion: a matter of phase control and amplitude modulation. Diabetologia. 2009;52:739-751. DOI: 10.1007/s00125-009-1314-y
  13. 13. Gembal M, Detimary P, Gilon P, Gao ZY, Henquin JC. Mechanisms by which glucose can control insulin release independently from its action on adenosine triphosphate-sensitive K+ channels in mouse B cells. The Journal of Clinical Investigation. 1993;91:871-880. DOI: 10.1172/jci116308
  14. 14. Komatsu M, Takei M, Ishii H, Sato Y. Glucose-stimulated insulin secretion: A newer perspective. Journal of diabetes investigation. 2013;4:511-516. DOI: 10.1111/jdi.12094
  15. 15. Pedersen MG, Tagliavini A, Henquin JC. Calcium signaling and secretory granule pool dynamics underlie biphasic insulin secretion and its amplification by glucose: experiments and modeling. American Journal of Physiology. Endocrinology and Metabolism. 2019;316:E475-e486. DOI: 10.1152/ajpendo.00380.2018
  16. 16. Daniel S, Noda M, Straub SG, Sharp GW. Identification of the docked granule pool responsible for the first phase of glucose-stimulated insulin secretion. Diabetes. 1999;48:1686-1690. DOI: 10.2337/diabetes.48.9.1686
  17. 17. Rorsman P, Renström E. Insulin granule dynamics in pancreatic beta cells. Diabetologia. 2003;46:1029-1045. DOI: 10.1007/s00125-003-1153-1
  18. 18. Nagamatsu S, Ohara-Imaizumi M, Nakamichi Y, Kikuta T, Nishiwaki C. Imaging docking and fusion of insulin granules induced by antidiabetes agents: sulfonylurea and glinide drugs preferentially mediate the fusion of newcomer, but not previously docked, insulin granules. Diabetes. 2006;55:2819-2825. DOI: 10.2337/db06-0105
  19. 19. Ohara-Imaizumi M, Fujiwara T, Nakamichi Y, Okamura T, Akimoto Y, Kawai J, et al. Imaging analysis reveals mechanistic differences between first- and second-phase insulin exocytosis. The Journal of cell biology. 2007;177:695-705. DOI: 10.1083/jcb.200608132
  20. 20. Kalwat MA, Thurmond DC. Signaling mechanisms of glucose-induced F-actin remodeling in pancreatic islet β cells. Experimental & Molecular Medicine. 2013;45:e37. DOI: 10.1038/emm.2013.73
  21. 21. Mourad NI, Nenquin M, Henquin JC. Metabolic amplifying pathway increases both phases of insulin secretion independently of beta-cell actin microfilaments. American Journal of Physiology. Cell Physiology. 2010;299:C389-C398. DOI: 10.1152/ajpcell.00138.2010
  22. 22. Wang Z, Thurmond DC. Mechanisms of biphasic insulin-granule exocytosis - roles of the cytoskeleton, small GTPases and SNARE proteins. Journal of Cell Science. 2009;122:893-903. DOI: 10.1242/jcs.034355
  23. 23. Mourad NI, Nenquin M, Henquin JC. cAMP-mediated and metabolic amplification of insulin secretion are distinct pathways sharing independence of β-cell microfilaments. Endocrinology. 2012;153:4644-4654. DOI: 10.1210/en.2012-1450
  24. 24. Mourad NI, Nenquin M, Henquin JC. Amplification of insulin secretion by acetylcholine or phorbol ester is independent of β-cell microfilaments and distinct from metabolic amplification. Molecular and Cellular Endocrinology. 2013;367:11-20. DOI: 10.1016/j.mce.2012.12.002
  25. 25. Shibasaki T, Takahashi H, Miki T, Sunaga Y, Matsumura K, Yamanaka M, et al. Essential role of Epac2/Rap1 signaling in regulation of insulin granule dynamics by cAMP. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:19333-19338. DOI: 10.1073/pnas.0707054104
  26. 26. Vakilian M, Tahamtani Y, Ghaedi K. A review on insulin trafficking and exocytosis. Gene. 2019;706:52-61. DOI: 10.1016/j.gene.2019.04.063
  27. 27. Rocheleau JV, Remedi MS, Granada B, Head WS, Koster JC, Nichols CG, et al. Critical role of gap junction coupled KATP channel activity for regulated insulin secretion. PLoS Biology. 2006;4:e26. DOI: 10.1371/journal.pbio.0040026
  28. 28. Idevall-Hagren O, Tengholm A. Metabolic regulation of calcium signaling in beta cells. Seminars in Cell & Developmental Biology. 2020;103:20-30. DOI: 10.1016/j.semcdb.2020.01.008
  29. 29. Johnston NR, Mitchell RK, Haythorne E, Pessoa MP, Semplici F, Ferrer J, et al. Beta Cell Hubs Dictate Pancreatic Islet Responses to Glucose. Cell Metabolism. 2016;24:389-401. DOI: 10.1016/j.cmet.2016.06.020
  30. 30. Ashcroft FM, Rorsman P. Diabetes Mellitus and the β Cell: The Last Ten Years. Cell. 2012;148:1160-1171. DOI: 10.1016/j.cell.2012.02.010
  31. 31. Maechler P. Mitochondrial function and insulin secretion. Molecular and Cellular Endocrinology. 2013. DOI: 10.1016/j.mce.2013.06.019
  32. 32. Prentki M, Matschinsky FM, Madiraju SRM. Metabolic signaling in fuel-induced insulin secretion. Cell Metabolism. 2013;18:162-185. DOI: 10.1016/j.cmet.2013.05.018
  33. 33. Rutter GA, Pullen TJ, Hodson DJ, Martinez-Sanchez A. Pancreatic β-cell identity, glucose sensing and the control of insulin secretion. Biochemical Journal. 2015;466:203-218. DOI: 10.1042/BJ20141384
  34. 34. Seino S, Sugawara K, Yokoi N, Takahashi H. β-Cell signalling and insulin secretagogues: A path for improved diabetes therapy. Diabetes, Obesity & Metabolism. 2017;19(Suppl 1):22-29. DOI: 10.1111/dom.12995
  35. 35. Shibasaki T, Takahashi T, Takahashi H, Seino S. Cooperation between cAMP signalling and sulfonylurea in insulin secretion. Diabetes, Obesity & Metabolism. 2014;16(Suppl 1):118-125. DOI: 10.1111/dom.12343
  36. 36. Seino S. Cell signalling in insulin secretion: the molecular targets of ATP, cAMP and sulfonylurea. Diabetologia. 2012;55:2096-2108. DOI: 10.1007/s00125-012-2562-9
  37. 37. Leloup C, Tourrel-Cuzin C, Magnan C, Karaca M, Castel J, Carneiro L, et al. Mitochondrial Reactive Oxygen Species Are Obligatory Signals for Glucose-Induced Insulin Secretion. Diabetes. 2009;58:673-681. DOI: 10.2337/db07-1056
  38. 38. Saadeh M, Ferrante TC, Kane A, Shirihai O, Corkey BE, Deeney JT. Reactive Oxygen Species Stimulate Insulin Secretion in Rat Pancreatic Islets: Studies Using Mono-Oleoyl-Glycerol. PLoS One. 2012;7:e30200. DOI: 10.1371/journal.pone.0030200
  39. 39. Rebelato E, Abdulkader F, Curi R, Carpinelli AR. Control of the Intracellular Redox State by Glucose Participates in the Insulin Secretion Mechanism. PLoS One. 2011;6:e24507. DOI: 10.1371/journal.pone.0024507
  40. 40. Pi J, Bai Y, Zhang Q , Wong V, Floering LM, Daniel K, et al. Reactive Oxygen Species as a Signal in Glucose-Stimulated Insulin Secretion. Diabetes. 2007;56:1783-1791. DOI: 10.2337/db06-1601
  41. 41. Morgan D, Rebelato E, Abdulkader F, Graciano MFR, Oliveira-Emilio HR, Hirata AE, et al. Association of NAD(P)H Oxidase with Glucose-Induced Insulin Secretion by Pancreatic β-Cells. Endocrinology. 2009;150:2197-2201. DOI: 10.1210/en.2008-1149
  42. 42. Imoto H, Sasaki N, Iwase M, Nakamura U, Oku M, Sonoki K, et al. Impaired Insulin Secretion by Diphenyleneiodium Associated with Perturbation of Cytosolic Ca 2+ Dynamics in Pancreatic β-Cells. Endocrinology. 2008;149:5391-5400. DOI: 10.1210/en.2008-0186
  43. 43. Syed I, Kyathanahalli CN, Kowluru A. Phagocyte-like NADPH oxidase generates ROS in INS 832/13 cells and rat islets: role of protein prenylation. American Journal of Physiology-Regulatory, Integrative and Comparative Physiology. 2011;300:R756-R762. DOI: 10.1152/ajpregu.00786.2010
  44. 44. Li N, Li B, Brun T, Deffert-Delbouille C, Mahiout Z, Daali Y, et al. NADPH Oxidase NOX2 Defines a New Antagonistic Role for Reactive Oxygen Species and cAMP/PKA in the Regulation of Insulin Secretion. Diabetes. 2012;61:2842-2850. DOI: 10.2337/db12-0009
  45. 45. Rorsman P, Eliasson L, Kanno T, Zhang Q , Gopel S. Electrophysiology of pancreatic β-cells in intact mouse islets of Langerhans. Progress in Biophysics and Molecular Biology. 2011;107:224-235. DOI: 10.1016/j.pbiomolbio.2011.06.009
  46. 46. Kakei M, Yoshida M, Dezaki K, Ito K, Yamada H, Funazaki S, et al. Glucose and GTP-binding protein-coupled receptor cooperatively regulate transient receptor potential-channels to stimulate insulin secretion [Review]. Endocrine Journal. 2016;63:867-876. DOI: 10.1507/endocrj.EJ16-0262
  47. 47. Sakaguchi R, Mori Y. Transient receptor potential (TRP) channels: Biosensors for redox environmental stimuli and cellular status. Free Radical Biology & Medicine. 2020;146:36-44. DOI: 10.1016/j.freeradbiomed.2019.10.415
  48. 48. Plecitá-Hlavatá L, Engstová H, Holendová B, Tauber J, Špaček T, Petrásková L, et al. Mitochondrial Superoxide Production Decreases on Glucose-Stimulated Insulin Secretion in Pancreatic β Cells Due to Decreasing Mitochondrial Matrix NADH/NAD(+) Ratio. Antioxidants & redox signaling. 2020. DOI: 10.1089/ars.2019.7800
  49. 49. Spégel P, Sharoyko VV, Goehring I, Danielsson AP, Malmgren S, Nagorny CL, et al. Time-resolved metabolomics analysis of β-cells implicates the pentose phosphate pathway in the control of insulin release. The Biochemical Journal. 2013;450:595-605. DOI: 10.1042/bj20121349
  50. 50. Bedard K, Krause K-H. The NOX Family of ROS-Generating NADPH Oxidases: Physiology and Pathophysiology. Physiological Reviews. 2007;87:245-313. DOI: 10.1152/physrev.00044.2005
  51. 51. Serrander L, Cartier L, Bedard K, Banfi B, Lardy B, Plastre O, et al. NOX4 activity is determined by mRNA levels and reveals a unique pattern of ROS generation. Biochemical Journal. 2007;406:105-114. DOI: 10.1042/BJ20061903
  52. 52. MacDonald PE. Signal integration at the level of ion channel and exocytotic function in pancreatic β-cells. American Journal of Physiology-Endocrinology and Metabolism. 2011;301:E1065-E1069. DOI: 10.1152/ajpendo.00426.2011
  53. 53. Leturque A, Brot-Laroche E, Le Gall M. GLUT2 mutations, translocation, and receptor function in diet sugar managing. American Journal of Physiology. Endocrinology and Metabolism. 2009;296:E985-E992. DOI: 10.1152/ajpendo.00004.2009
  54. 54. Kaminski MT, Lenzen S, Baltrusch S. Real-time analysis of intracellular glucose and calcium in pancreatic beta cells by fluorescence microscopy. Biochimica et Biophysica Acta. 2012;1823:1697-1707. DOI: 10.1016/j.bbamcr.2012.06.022
  55. 55. Park JH, Kim SJ, Park SH, Son DG, Bae JH, Kim HK, et al. Glucagon-like peptide-1 enhances glucokinase activity in pancreatic beta-cells through the association of Epac2 with Rim2 and Rab3A. Endocrinology. 2012;153:574-582. DOI: 10.1210/en.2011-0259
  56. 56. Matschinsky FM, Wilson DF. The Central Role of Glucokinase in Glucose Homeostasis: A Perspective 50 Years After Demonstrating the Presence of the Enzyme in Islets of Langerhans. Frontiers in Physiology. 2019;10:148. DOI: 10.3389/fphys.2019.00148
  57. 57. Schuit F, De Vos A, Farfari S, Moens K, Pipeleers D, Brun T, et al. Metabolic fate of glucose in purified islet cells. Glucose-regulated anaplerosis in beta cells. The Journal of Biological Chemistry. 1997;272:18572-18579
  58. 58. Zhang Z, Liew CW, Handy DE, Zhang Y, Leopold JA, Hu J, et al. High glucose inhibits glucose-6-phosphate dehydrogenase, leading to increased oxidative stress and beta-cell apoptosis. FASEB Journal : Official Publication of the Federation of American Societies for Experimental Biology. 2010;24:1497-1505. DOI: 10.1096/fj.09-136572
  59. 59. Huang M, Joseph JW. Metabolomic analysis of pancreatic β-cell insulin release in response to glucose. Islets. 2012;4:210-222. DOI: 10.4161/isl.20141
  60. 60. Goehring I, Sauter NS, Catchpole G, Assmann A, Shu L, Zien KS, et al. Identification of an intracellular metabolic signature impairing beta cell function in the rat beta cell line INS-1E and human islets. Diabetologia. 2011;54:2584-2594. DOI: 10.1007/s00125-011-2249-7
  61. 61. Ammon HP, Steinke J. 6-Amnionicotinamide (6-AN) as a diabetogenic agent. In vitro and in vivo studies in the rat. Diabetes. 1972;21:143-148. DOI: 10.2337/diab.21.3.143
  62. 62. Verspohl EJ, Händel M, Ammon HP. Pentosephosphate shunt activity of rat pancreatic islets: its dependence on glucose concentration. Endocrinology. 1979;105:1269-1274. DOI: 10.1210/endo-105-5-1269
  63. 63. Monte Alegre S, Saad ST, Delatre E, Saad MJ. Insulin secretion in patients deficient in glucose-6-phosphate dehydrogenase. Hormone and Metabolic Research. 1991;23:171-173. DOI: 10.1055/s-2007-1003644
  64. 64. Akhmedov D, De Marchi U, Wollheim CB, Wiederkehr A. Pyruvate dehydrogenase E1α phosphorylation is induced by glucose but does not control metabolism-secretion coupling in INS-1E clonal β-cells. Biochimica et Biophysica Acta. 2012;1823:1815-1824. DOI: 10.1016/j.bbamcr.2012.07.005
  65. 65. Rutter GA, Hodson DJ, Chabosseau P, Haythorne E, Pullen TJ, Leclerc I. Local and regional control of calcium dynamics in the pancreatic islet. Diabetes, Obesity & Metabolism. 2017;19(Suppl 1):30-41. DOI: 10.1111/dom.12990
  66. 66. Alves TC, Pongratz RL, Zhao X, Yarborough O, Sereda S, Shirihai O, et al. Integrated, Step-Wise, Mass-Isotopomeric Flux Analysis of the TCA Cycle. Cell Metabolism. 2015;22:936-947. DOI: 10.1016/j.cmet.2015.08.021
  67. 67. Kahancová A, Sklenář F, Ježek P, Dlasková A. Regulation of glucose-stimulated insulin secretion by ATPase Inhibitory Factor 1 (IF1). FEBS Letters. 2018;592:999-1009. DOI: 10.1002/1873-3468.12991
  68. 68. Kahancová A, Sklenář F, Ježek P, Dlasková A. Overexpression of native IF1 downregulates glucose-stimulated insulin secretion by pancreatic INS-1E cells. Scientific Reports. 2020;10:1551. DOI: 10.1038/s41598-020-58411-x
  69. 69. Yang SN, Shi Y, Yang G, Li Y, Yu J, Berggren PO. Ionic mechanisms in pancreatic β cell signaling. Cellular and Molecular Life Sciences. 2014;71:4149-4177. DOI: 10.1007/s00018-014-1680-6
  70. 70. Drews G, Krippeit-Drews P, Düfer M. Electrophysiology of Islet Cells. Advances in Experimental Medicine and Biology. 2010;654:115-163
  71. 71. Bennett K, James C, Hussain K. Pancreatic β-cell KATP channels: Hypoglycaemia and hyperglycaemia. Reviews in Endocrine and Metabolic Disorders. 2010;11:157-163. DOI: 10.1007/s11154-010-9144-2
  72. 72. Szollosi A, Nenquin M, Henquin J. Pharmacological stimulation and inhibition of insulin secretion in mouse islets lacking ATP-sensitive K+ channels. British Journal of Pharmacology. 2010;159:669-677. DOI: 10.1111/j.1476-5381.2009.00588.x
  73. 73. Soty M, Visa M, Soriano S, del Carmen Carmona M, Nadal Á, Novials A. Involvement of ATP-sensitive Potassium (K ATP ) Channels in the Loss of Beta-cell Function Induced by Human Islet Amyloid Polypeptide. Journal of Biological Chemistry. 2011;286:40857-40866. DOI: 10.1074/jbc.M111.232801
  74. 74. Rorsman P, Braun M, Zhang Q . Regulation of calcium in pancreatic α- and β-cells in health and disease. Cell Calcium. 2012;51:300-308. DOI: 10.1016/j.ceca.2011.11.006
  75. 75. Rorsman P, Braun M. Regulation of insulin secretion in human pancreatic islets. Annual Review of Physiology. 2013;75:155-179. DOI: 10.1146/annurev-physiol-030212-183754
  76. 76. Smith PA, Ashcroft FM, Rorsman P. Simultaneous recordings of glucose dependent electrical activity and ATP-regulated K(+)-currents in isolated mouse pancreatic beta-cells. FEBS Letters. 1990;261:187-190. DOI: 10.1016/0014-5793(90)80667-8
  77. 77. Best L. Glucose-induced electrical activity in rat pancreatic beta-cells: dependence on intracellular chloride concentration. The Journal of Physiology. 2005;568:137-144. DOI: 10.1113/jphysiol.2005.093740
  78. 78. Stuhlmann T, Planells-Cases R, Jentsch TJ. LRRC8/VRAC anion channels enhance β-cell glucose sensing and insulin secretion. Nature Communications. 2018;9:1974. DOI: 10.1038/s41467-018-04353-y
  79. 79. Di Fulvio M, Aguilar-Bryan L. Chloride transporters and channels in β-cell physiology: revisiting a 40-year-old model. Biochemical Society Transactions. 2019;47:1843-1855. DOI: 10.1042/bst20190513
  80. 80. Colsoul B, Schraenen A, Lemaire K, Quintens R, Van Lommel L, Segal A, et al. Loss of high-frequency glucose-induced Ca2+ oscillations in pancreatic islets correlates with impaired glucose tolerance in Trpm5−/− mice. Proceedings of the National Academy of Sciences of the United States of America. 2010;107:5208-5213. DOI: 10.1073/pnas.0913107107
  81. 81. Gilon P, Ravier MA, Jonas JC, Henquin JC. Control mechanisms of the oscillations of insulin secretion in vitro and in vivo. Diabetes. 2002;51(Suppl 1):S144-S151. DOI: 10.2337/diabetes.51.2007.s144
  82. 82. Beauvois MC, Merezak C, Jonas JC, Ravier MA, Henquin JC, Gilon P. Glucose-induced mixed [Ca2+]c oscillations in mouse beta-cells are controlled by the membrane potential and the SERCA3 Ca2+-ATPase of the endoplasmic reticulum. American Journal of Physiology. Cell Physiology. 2006;290:C1503-C1511. DOI: 10.1152/ajpcell.00400.2005
  83. 83. Sabourin J, Allagnat F. Store-operated Ca2+ entry: a key component of the insulin secretion machinery. Journal of Molecular Endocrinology. 2016;57:F35-f39. DOI: 10.1530/jme-16-0106
  84. 84. Sabourin J, Le Gal L, Saurwein L, Haefliger JA, Raddatz E, Allagnat F. Store-operated Ca2+ Entry Mediated by Orai1 and TRPC1 Participates to Insulin Secretion in Rat β-Cells. The Journal of Biological Chemistry. 2015;290:30530-30539. DOI: 10.1074/jbc.M115.682583
  85. 85. Rorsman P, Trube G. Calcium and delayed potassium currents in mouse pancreatic beta-cells under voltage-clamp conditions. The Journal of Physiology. 1986;374:531-550. DOI: 10.1113/jphysiol.1986.sp016096
  86. 86. Jacobson DA, Kuznetsov A, Lopez JP, Kash S, Ammälä CE, Philipson LH. Kv2.1 ablation alters glucose-induced islet electrical activity, enhancing insulin secretion. Cell Metabolism. 2007;6:229-235. DOI: 10.1016/j.cmet.2007.07.010
  87. 87. Rebelato E, Santos LR, Carpinelli AR, Rorsman P, Abdulkader F. Short-term high glucose culture potentiates pancreatic beta cell function. Scientific Reports. 2018;8:13061. DOI: 10.1038/s41598-018-31325-5
  88. 88. Li N, Wu JX, Ding D, Cheng J, Gao N, Chen L. Structure of a Pancreatic ATP-Sensitive Potassium Channel. Cell. 2017;168:101-110.e110. DOI: 10.1016/j.cell.2016.12.028
  89. 89. Martin GM, Yoshioka C, Rex EA, Fay JF, Xie Q , Whorton MR, et al. Cryo-EM structure of the ATP-sensitive potassium channel illuminates mechanisms of assembly and gating. eLife. 2017;6. DOI: 10.7554/eLife.24149
  90. 90. Mikhailov MV, Campbell JD, de Wet H, Shimomura K, Zadek B, Collins RF, et al. 3-D structural and functional characterization of the purified KATP channel complex Kir6.2-SUR1. EMBO J. 2005;24:4166-4175. DOI: 10.1038/sj.emboj.7600877
  91. 91. Nichols CG. KATP channels as molecular sensors of cellular metabolism. Nature. 2006;440:470-476. DOI: 10.1038/nature04711
  92. 92. Yang HQ , Martinez-Ortiz W, Hwang J, Fan X, Cardozo TJ, Coetzee WA. Palmitoylation of the K(ATP) channel Kir6.2 subunit promotes channel opening by regulating PIP(2) sensitivity. Proceedings of the National Academy of Sciences of the United States of America. 2020;117:10593-10602. DOI: 10.1073/pnas.1918088117
  93. 93. Shyng SL, Nichols CG. Membrane phospholipid control of nucleotide sensitivity of KATP channels. Science. 1998;282:1138-1141. DOI: 10.1126/science.282.5391.1138
  94. 94. Baukrowitz, T.; Schulte, U.; Oliver, D.; Herlitze, S.; Krauter, T.; Tucker, S.J.; Ruppersberg, J.P.; Fakler, B. PIP2 and PIP as determinants for ATP inhibition of KATP channels. Science 1998, 282, 1141-1144, doi:10.1126/science.282.5391.1141
  95. 95. Shyng S, Ferrigni T, Nichols CG. Regulation of KATP channel activity by diazoxide and MgADP. Distinct functions of the two nucleotide binding folds of the sulfonylurea receptor. J Gen Physiol. 1997;110:643-654. DOI: 10.1085/jgp.110.6.643
  96. 96. Lin YF, Jan YN, Jan LY. Regulation of ATP-sensitive potassium channel function by protein kinase A-mediated phosphorylation in transfected HEK293 cells. The EMBO Journal. 2000;19:942-955. DOI: 10.1093/emboj/19.5.942
  97. 97. Béguin P, Nagashima K, Nishimura M, Gonoi T, Seino S. PKA-mediated phosphorylation of the human K(ATP) channel: separate roles of Kir6.2 and SUR1 subunit phosphorylation. Embo j. 1999;18:4722-4732. DOI: 10.1093/emboj/18.17.4722
  98. 98. Kline CF, Wright PJ, Koval OM, Zmuda EJ, Johnson BL, Anderson ME, et al. βIV-Spectrin and CaMKII facilitate Kir6.2 regulation in pancreatic beta cells. Proceedings of the National Academy of Sciences of the United States of America. 2013;110:17576-17581. DOI: 10.1073/pnas.1314195110
  99. 99. Ashcroft FM, Harrison DE, Ashcroft SJ. Glucose induces closure of single potassium channels in isolated rat pancreatic beta-cells. Nature. 1984;312:446-448. DOI: 10.1038/312446a0
  100. 100. Yasui S, Mawatari K, Morizumi R, Furukawa H, Shimohata T, Harada N, et al. Hydrogen peroxide inhibits insulin-induced ATP-sensitive potassium channel activation independent of insulin signaling pathway in cultured vascular smooth muscle cells. The journal of medical investigation: JMI. 2012;59:36-44
  101. 101. Finol-Urdaneta RK, Remedi MS, Raasch W, Becker S, Clark RB, Strüver N, et al. Block of Kv1.7 potassium currents increases glucose-stimulated insulin secretion. EMBO Molecular Medicine. 2012;4:424-434. DOI: 10.1002/emmm.201200218
  102. 102. MacDonald PE, Salapatek AM, Wheeler MB. Temperature and redox state dependence of native Kv2.1 currents in rat pancreatic beta-cells. The Journal of Physiology. 2003;546:647-653. DOI: 10.1113/jphysiol.2002.035709
  103. 103. Mittal M, Gu XQ , Pak O, Pamenter ME, Haag D, Fuchs DB, et al. Hypoxia induces Kv channel current inhibition by increased NADPH oxidase-derived reactive oxygen species. Free Radical Biology & Medicine. 2012;52:1033-1042. DOI: 10.1016/j.freeradbiomed.2011.12.004
  104. 104. Llanos P, Contreras-Ferrat A, Barrientos G, Valencia M, Mears D, Hidalgo C. Glucose-Dependent Insulin Secretion in Pancreatic β-Cell Islets from Male Rats Requires Ca2+ Release via ROS-Stimulated Ryanodine Receptors. PLoS One. 2015;10:e0129238. DOI: 10.1371/journal.pone.0129238
  105. 105. Kashio M, Tominaga M. Redox Signal-mediated Enhancement of the Temperature Sensitivity of Transient Receptor Potential Melastatin 2 (TRPM2) Elevates Glucose-induced Insulin Secretion from Pancreatic Islets. Journal of Biological Chemistry. 2015;290:12435-12442. DOI: 10.1074/jbc.M115.649913
  106. 106. Yosida M, Dezaki K, Uchida K, Kodera S, Lam NV, Ito K, et al. Involvement of cAMP/EPAC/TRPM2 activation in glucose- and incretin-induced insulin secretion. Diabetes. 2014;63:3394-3403. DOI: 10.2337/db13-1868
  107. 107. Müller TD, Finan B, Bloom SR, D’Alessio D, Drucker DJ, Flatt PR, et al. Glucagon-like peptide 1 (GLP-1). Molecular metabolism. 2019;30:72-130. DOI: 10.1016/j.molmet.2019.09.010
  108. 108. Furman B, Ong WK, Pyne NJ. Cyclic AMP signaling in pancreatic islets. Advances in Experimental Medicine and Biology. 2010;654:281-304. DOI: 10.1007/978-90-481-3271-3_13
  109. 109. Lefkimmiatis K, Zaccolo M. cAMP signaling in subcellular compartments. Pharmacology & Therapeutics. 2014;143:295-304. DOI: 10.1016/j.pharmthera.2014.03.008
  110. 110. Berridge MJ. The Inositol Trisphosphate/Calcium Signaling Pathway in Health and Disease. Physiological Reviews. 2016;96:1261-1296. DOI: 10.1152/physrev.00006.2016
  111. 111. Husted AS, Trauelsen M, Rudenko O, Hjorth SA, Schwartz TW. GPCR-Mediated Signaling of Metabolites. Cell Metabolism. 2017;25:777-796. DOI: 10.1016/j.cmet.2017.03.008
  112. 112. Salloum G, Jaafar L, El-Sibai M. Rho A and Rac1: Antagonists moving forward. Tissue & Cell. 2020;65:101364. DOI: 10.1016/j.tice.2020.101364
  113. 113. Dalle S, Ravier MA, Bertrand G. Emerging roles for β-arrestin-1 in the control of the pancreatic β-cell function and mass: new therapeutic strategies and consequences for drug screening. Cellular Signalling. 2011;23:522-528. DOI: 10.1016/j.cellsig.2010.09.014
  114. 114. Taylor SS, Ilouz R, Zhang P, Kornev AP. Assembly of allosteric macromolecular switches: lessons from PKA. Nature Reviews. Molecular Cell Biology. 2012;13:646-658. DOI: 10.1038/nrm3432
  115. 115. Zhang F, Zhang L, Qi Y, Xu H. Mitochondrial cAMP signaling. Cellular and Molecular Life Sciences. 2016;73:4577-4590. DOI: 10.1007/s00018-016-2282-2
  116. 116. Ould Amer Y, Hebert-Chatelain E. Mitochondrial cAMP-PKA signaling: What do we really know? Biochimica et Biophysica Acta - Bioenergetics. 2018;1859:868-877. DOI: 10.1016/j.bbabio.2018.04.005
  117. 117. Kang G, Leech CA, Chepurny OG, Coetzee WA, Holz GG. Role of the cAMP sensor Epac as a determinant of KATP channel ATP sensitivity in human pancreatic beta-cells and rat INS-1 cells. The Journal of Physiology. 2008;586:1307-1319. DOI: 10.1113/jphysiol.2007.143818
  118. 118. de Rooij J, Zwartkruis FJ, Verheijen MH, Cool RH, Nijman SM, Wittinghofer A, et al. Epac is a Rap1 guanine-nucleotide-exchange factor directly activated by cyclic AMP. Nature. 1998;396:474-477. DOI: 10.1038/24884
  119. 119. Gloerich M, Bos JL. Epac: defining a new mechanism for cAMP action. Annual Review of Pharmacology and Toxicology. 2010;50:355-375. DOI: 10.1146/annurev.pharmtox.010909.105714
  120. 120. Härndahl L, Jing XJ, Ivarsson R, Degerman E, Ahrén B, Manganiello VC, et al. Important role of phosphodiesterase 3B for the stimulatory action of cAMP on pancreatic beta-cell exocytosis and release of insulin. The Journal of Biological Chemistry. 2002;277:37446-37455. DOI: 10.1074/jbc.M205401200
  121. 121. Bünemann M, Gerhardstein BL, Gao T, Hosey MM. Functional regulation of L-type calcium channels via protein kinase A-mediated phosphorylation of the beta(2) subunit. The Journal of Biological Chemistry. 1999;274:33851-33854. DOI: 10.1074/jbc.274.48.33851
  122. 122. MacDonald PE, Wang X, Xia F, El-kholy W, Targonsky ED, Tsushima RG, et al. Antagonism of rat beta-cell voltage-dependent K+ currents by exendin 4 requires dual activation of the cAMP/protein kinase A and phosphatidylinositol 3-kinase signaling pathways. The Journal of Biological Chemistry. 2003;278:52446-52453. DOI: 10.1074/jbc.M307612200
  123. 123. Song WJ, Seshadri M, Ashraf U, Mdluli T, Mondal P, Keil M, et al. Snapin mediates incretin action and augments glucose-dependent insulin secretion. Cell Metabolism. 2011;13:308-319. DOI: 10.1016/j.cmet.2011.02.002
  124. 124. Somanath S, Partridge CJ, Marshall C, Rowe T, Turner MD. Snapin mediates insulin secretory granule docking, but not trans-SNARE complex formation. Biochemical and Biophysical Research Communications. 2016;473:403-407. DOI: 10.1016/j.bbrc.2016.02.123
  125. 125. Holz GG. Epac: A new cAMP-binding protein in support of glucagon-like peptide-1 receptor-mediated signal transduction in the pancreatic beta-cell. Diabetes. 2004;53:5-13. DOI: 10.2337/diabetes.53.1.5
  126. 126. Holz GG, Leech CA, Heller RS, Castonguay M, Habener JF. cAMP-dependent mobilization of intracellular Ca2+ stores by activation of ryanodine receptors in pancreatic beta-cells. A Ca2+ signaling system stimulated by the insulinotropic hormone glucagon-like peptide-1-(7-37). The Journal of Biological Chemistry. 1999;274:14147-14156. DOI: 10.1074/jbc.274.20.14147
  127. 127. Gilon P, Chae HY, Rutter GA, Ravier MA. Calcium signaling in pancreatic β-cells in health and in Type 2 diabetes. Cell Calcium. 2014;56:340-361. DOI: 10.1016/j.ceca.2014.09.001
  128. 128. Kang G, Chepurny OG, Holz GG. cAMP-regulated guanine nucleotide exchange factor II (Epac2) mediates Ca2+−induced Ca2+ release in INS-1 pancreatic beta-cells. The Journal of Physiology. 2001;536:375-385. DOI: 10.1111/j.1469-7793.2001.0375c.xd
  129. 129. Ozaki N, Shibasaki T, Kashima Y, Miki T, Takahashi K, Ueno H, et al. cAMP-GEFII is a direct target of cAMP in regulated exocytosis. Nature Cell Biology. 2000;2:805-811. DOI: 10.1038/35041046
  130. 130. Kashima Y, Miki T, Shibasaki T, Ozaki N, Miyazaki M, Yano H, et al. Critical role of cAMP-GEFII--Rim2 complex in incretin-potentiated insulin secretion. The Journal of Biological Chemistry. 2001;276:46046-46053. DOI: 10.1074/jbc.M108378200
  131. 131. Yasuda T, Shibasaki T, Minami K, Takahashi H, Mizoguchi A, Uriu Y, et al. Rim2alpha determines docking and priming states in insulin granule exocytosis. Cell Metabolism. 2010;12:117-129. DOI: 10.1016/j.cmet.2010.05.017
  132. 132. Acin-Perez R, Russwurm M, Günnewig K, Gertz M, Zoidl G, Ramos L, et al. A phosphodiesterase 2A isoform localized to mitochondria regulates respiration. The Journal of Biological Chemistry. 2011;286:30423-30432. DOI: 10.1074/jbc.M111.266379
  133. 133. Zhang F, Qi Y, Zhou K, Zhang G, Linask K, Xu H. The cAMP phosphodiesterase Prune localizes to the mitochondrial matrix and promotes mtDNA replication by stabilizing TFAM. EMBO Reports. 2015;16:520-527. DOI: 10.15252/embr.201439636
  134. 134. DiPilato LM, Cheng X, Zhang J. Fluorescent indicators of cAMP and Epac activation reveal differential dynamics of cAMP signaling within discrete subcellular compartments. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:16513-16518. DOI: 10.1073/pnas.0405973101
  135. 135. Di Benedetto G, Scalzotto E, Mongillo M, Pozzan T. Mitochondrial Ca2+ uptake induces cyclic AMP generation in the matrix and modulates organelle ATP levels. Cell Metabolism. 2013;17:965-975. DOI: 10.1016/j.cmet.2013.05.003
  136. 136. Chen Y, Cann MJ, Litvin TN, Iourgenko V, Sinclair ML, Levin LR, et al. Soluble adenylyl cyclase as an evolutionarily conserved bicarbonate sensor. Science. 2000;289:625-628. DOI: 10.1126/science.289.5479.625
  137. 137. Lefkimmiatis K, Leronni D, Hofer AM. The inner and outer compartments of mitochondria are sites of distinct cAMP/PKA signaling dynamics. The Journal of Cell Biology. 2013;202:453-462. DOI: 10.1083/jcb.201303159
  138. 138. Agnes RS, Jernigan F, Shell JR, Sharma V, Lawrence DS. Suborganelle sensing of mitochondrial cAMP-dependent protein kinase activity. Journal of the American Chemical Society. 2010;132:6075-6080. DOI: 10.1021/ja909652q
  139. 139. De Rasmo D, Micelli L, Santeramo A, Signorile A, Lattanzio P, Papa S. cAMP regulates the functional activity, coupling efficiency and structural organization of mammalian FOF1 ATP synthase. Biochimica et Biophysica Acta. 2016;1857:350-358. DOI: 10.1016/j.bbabio.2016.01.006
  140. 140. García-Bermúdez J, Sánchez-Aragó M, Soldevilla B, Del Arco A, Nuevo-Tapioles C, Cuezva JM. PKA Phosphorylates the ATPase Inhibitory Factor 1 and Inactivates Its Capacity to Bind and Inhibit the Mitochondrial H(+)-ATP Synthase. Cell Reports. 2015;12:2143-2155. DOI: 10.1016/j.celrep.2015.08.052
  141. 141. Srinivasan S, Spear J, Chandran K, Joseph J, Kalyanaraman B, Avadhani NG. Oxidative stress induced mitochondrial protein kinase A mediates cytochrome c oxidase dysfunction. PLoS One. 2013;8:e77129. DOI: 10.1371/journal.pone.0077129
  142. 142. Rosca M, Minkler P, Hoppel CL. Cardiac mitochondria in heart failure: normal cardiolipin profile and increased threonine phosphorylation of complex IV. Biochimica et Biophysica Acta. 2011;1807:1373-1382. DOI: 10.1016/j.bbabio.2011.02.003
  143. 143. Barker CJ, Berggren PO. New horizons in cellular regulation by inositol polyphosphates: insights from the pancreatic β-cell. Pharmacological Reviews. 2013;65:641-669. DOI: 10.1124/pr.112.006775
  144. 144. Shigeto M, Ramracheya R, Tarasov AI, Cha CY, Chibalina MV, Hastoy B, et al. GLP-1 stimulates insulin secretion by PKC-dependent TRPM4 and TRPM5 activation. The Journal of Clinical Investigation. 2015;125:4714-4728. DOI: 10.1172/jci81975
  145. 145. Santo-Domingo J, Chareyron I, Dayon L, Núñez Galindo A, Cominetti O, Pilar Giner Giménez M, et al. Coordinated activation of mitochondrial respiration and exocytosis mediated by PKC signaling in pancreatic β cells. FASEB Journal : Official Publication of the Federation of American Societies for Experimental Biology. 2017;31:1028-1045. DOI: 10.1096/fj.201600837R
  146. 146. Elrick H, Stimmler L, Hlad CJ Jr, Arai Y. Plasma insulin response to oral and intravenous glucose administration. The Journal of Clinical Endocrinology and Metabolism. 1964;24:1076-1082. DOI: 10.1210/jcem-24-10-1076
  147. 147. Ebert R, Unger H, Creutzfeldt W. Preservation of incretin activity after removal of gastric inhibitory polypeptide (GIP) from rat gut extracts by immunoadsorption. Diabetologia. 1983;24:449-454. DOI: 10.1007/bf00257346
  148. 148. Scrocchi LA, Brown TJ, MaClusky N, Brubaker PL, Auerbach AB, Joyner AL, et al. Glucose intolerance but normal satiety in mice with a null mutation in the glucagon-like peptide 1 receptor gene. Nature Medicine. 1996;2:1254-1258. DOI: 10.1038/nm1196-1254
  149. 149. Scrocchi LA, Marshall BA, Cook SM, Brubaker PL, Drucker DJ. Identification of glucagon-like peptide 1 (GLP-1) actions essential for glucose homeostasis in mice with disruption of GLP-1 receptor signaling. Diabetes. 1998;47:632-639. DOI: 10.2337/diabetes.47.4.632
  150. 150. Drucker DJ, Philippe J, Mojsov S, Chick WL, Habener JF. Glucagon-like peptide I stimulates insulin gene expression and increases cyclic AMP levels in a rat islet cell line. Proceedings of the National Academy of Sciences of the United States of America. 1987;84:3434-3438. DOI: 10.1073/pnas.84.10.3434
  151. 151. Weir GC, Mojsov S, Hendrick GK, Habener JF. Glucagonlike peptide I (7-37) actions on endocrine pancreas. Diabetes. 1989;38:338-342. DOI: 10.2337/diab.38.3.338
  152. 152. Teraoku H, Lenzen S. Dynamics of Insulin Secretion from EndoC-βH1 β-Cell Pseudoislets in Response to Glucose and Other Nutrient and Nonnutrient Secretagogues. Journal Diabetes Research. 2017;2017:2309630. DOI: 10.1155/2017/2309630
  153. 153. Graaf Cd, Donnelly D, Wootten D, Lau J, Sexton PM, Miller LJ, et al. Glucagon-Like Peptide-1 and Its Class B G Protein-Coupled Receptors: A Long March to Therapeutic Successes. Pharmacological Reviews. 2016;68:954-1013. DOI: 10.1124/pr.115.011395
  154. 154. Hjøllund KR, Deacon CF, Holst JJ. Dipeptidyl peptidase-4 inhibition increases portal concentrations of intact glucagon-like peptide-1 (GLP-1) to a greater extent than peripheral concentrations in anaesthetised pigs. Diabetologia. 2011;54:2206-2208. DOI: 10.1007/s00125-011-2168-7
  155. 155. Kuhre RE, Wewer Albrechtsen NJ, Hartmann B, Deacon CF, Holst JJ. Measurement of the incretin hormones: glucagon-like peptide-1 and glucose-dependent insulinotropic peptide. Journal of Diabetes and its Complications. 2015;29:445-450. DOI: 10.1016/j.jdiacomp.2014.12.006
  156. 156. Herrmann C, Göke R, Richter G, Fehmann HC, Arnold R, Göke B. Glucagon-like peptide-1 and glucose-dependent insulin-releasing polypeptide plasma levels in response to nutrients. Digestion. 1995;56:117-126. DOI: 10.1159/000201231
  157. 157. Moran BM, Abdel-Wahab YH, Flatt PR, McKillop AM. Activation of GPR119 by fatty acid agonists augments insulin release from clonal β-cells and isolated pancreatic islets and improves glucose tolerance in mice. Biological Chemistry. 2014;395:453-464. DOI: 10.1515/hsz-2013-0255
  158. 158. Moon MJ, Park S, Kim DK, Cho EB, Hwang JI, Vaudry H, et al. Structural and molecular conservation of glucagon-like Peptide-1 and its receptor confers selective ligand-receptor interaction. Frontiers in Endocrinology. 2012;3:141. DOI: 10.3389/fendo.2012.00141
  159. 159. Wootten D, Reynolds CA, Smith KJ, Mobarec JC, Koole C, Savage EE, et al. The Extracellular Surface of the GLP-1 Receptor Is a Molecular Trigger for Biased Agonism. Cell. 2016;165:1632-1643. DOI: 10.1016/j.cell.2016.05.023
  160. 160. Sonoda N, Imamura T, Yoshizaki T, Babendure JL, Lu JC, Olefsky JM. Beta-Arrestin-1 mediates glucagon-like peptide-1 signaling to insulin secretion in cultured pancreatic beta cells. Proceedings of the National Academy of Sciences of the United States of America. 2008;105:6614-6619. DOI: 10.1073/pnas.0710402105
  161. 161. Montrose-Rafizadeh C, Avdonin P, Garant MJ, Rodgers BD, Kole S, Yang H, et al. Pancreatic glucagon-like peptide-1 receptor couples to multiple G proteins and activates mitogen-activated protein kinase pathways in Chinese hamster ovary cells. Endocrinology. 1999;140:1132-1140. DOI: 10.1210/endo.140.3.6550
  162. 162. Light PE, Manning Fox JE, Riedel MJ, Wheeler MB. Glucagon-like peptide-1 inhibits pancreatic ATP-sensitive potassium channels via a protein kinase A- and ADP-dependent mechanism. Molecular Endocrinology. 2002;16:2135-2144. DOI: 10.1210/me.2002-0084
  163. 163. Kang G, Joseph JW, Chepurny OG, Monaco M, Wheeler MB, Bos JL, et al. Epac-selective cAMP analog 8-pCPT-2'-O-Me-cAMP as a stimulus for Ca2+−induced Ca2+ release and exocytosis in pancreatic beta-cells. The Journal of Biological Chemistry. 2003;278:8279-8285. DOI: 10.1074/jbc.M211682200
  164. 164. Thompson A, Kanamarlapudi V. Agonist-induced internalisation of the glucagon-like peptide-1 receptor is mediated by the Gαq pathway. Biochemical Pharmacology. 2015;93:72-84. DOI: 10.1016/j.bcp.2014.10.015
  165. 165. MacDonald PE, Salapatek AM, Wheeler MB. Glucagon-like peptide-1 receptor activation antagonizes voltage-dependent repolarizing K(+) currents in beta-cells: a possible glucose-dependent insulinotropic mechanism. Diabetes. 2002;51(Suppl 3):S443-S447. DOI: 10.2337/diabetes.51.2007.s443
  166. 166. Panten U, Früh E, Reckers K, Rustenbeck I. Acute metabolic amplification of insulin secretion in mouse islets: Role of cytosolic acetyl-CoA. Metabolism: clinical and experimental. 2016;65:1225-1229. DOI: 10.1016/j.metabol.2016.05.001
  167. 167. Panten U, Willenborg M, Schumacher K, Hamada A, Ghaly H, Rustenbeck I. Acute metabolic amplification of insulin secretion in mouse islets is mediated by mitochondrial export of metabolites, but not by mitochondrial energy generation. Metabolism: clinical and experimental. 2013;62:1375-1386. DOI: 10.1016/j.metabol.2013.05.006
  168. 168. Hull J, Hindy ME, Kehoe PG, Chalmers K, Love S, Conway ME. Distribution of the branched chain aminotransferase proteins in the human brain and their role in glutamate regulation. Journal of Neurochemistry. 2012;123:997-1009. DOI: 10.1111/jnc.12044
  169. 169. Brand MD. Mitochondrial generation of superoxide and hydrogen peroxide as the source of mitochondrial redox signaling. Free Radical Biology and Medicine. 2016;100:14-31. DOI: 10.1016/j.freeradbiomed.2016.04.001
  170. 170. Nyrén R, Chang CL, Lindström P, Barmina A, Vorrsjö E, Ali Y, et al. Localization of lipoprotein lipase and GPIHBP1 in mouse pancreas: effects of diet and leptin deficiency. BMC Physiology. 2012;12:14. DOI: 10.1186/1472-6793-12-14
  171. 171. Winzell MS, Ström K, Holm C, Ahrén B. Glucose-stimulated insulin secretion correlates with beta-cell lipolysis. Nutrition, Metabolism, and Cardiovascular Diseases. 2006;16(Suppl 1):S11-S16. DOI: 10.1016/j.numecd.2005.11.006
  172. 172. Cruz WS, Kwon G, Marshall CA, McDaniel ML, Semenkovich CF. Glucose and insulin stimulate heparin-releasable lipoprotein lipase activity in mouse islets and INS-1 cells. A potential link between insulin resistance and beta-cell dysfunction. The Journal of Biological Chemistry. 2001;276:12162-12168. DOI: 10.1074/jbc.M010707200
  173. 173. Marshall BA, Tordjman K, Host HH, Ensor NJ, Kwon G, Marshall CA, et al. Relative hypoglycemia and hyperinsulinemia in mice with heterozygous lipoprotein lipase (LPL) deficiency. Islet LPL regulates insulin secretion. The Journal of biological chemistry. 1999;274:27426-27432. DOI: 10.1074/jbc.274.39.27426
  174. 174. Itoh K, Moriguchi R, Yamada Y, Fujita M, Yamato T, Oumi M, et al. High saturated fatty acid intake induces insulin secretion by elevating gastric inhibitory polypeptide levels in healthy individuals. Nutrition Research. 2014;34:653-660. DOI: 10.1016/j.nutres.2014.07.013
  175. 175. Ježek J, Dlasková A, Zelenka J, Jabůrek M, Ježek P. H2O2-Activated Mitochondrial Phospholipase iPLA2γ Prevents Lipotoxic Oxidative Stress in Synergy with UCP2, Amplifies Signaling via G-Protein–Coupled Receptor GPR40, and Regulates Insulin Secretion in Pancreatic β-Cells. Antioxidants & Redox Signaling. 2015;23:958-972. DOI: 10.1089/ars.2014.6195
  176. 176. Yamada H, Yoshida M, Ito K, Dezaki K, Yada T, Ishikawa SE, et al. Potentiation of Glucose-stimulated Insulin Secretion by the GPR40-PLC-TRPC Pathway in Pancreatic β-Cells. Scientific Reports. 2016;6:25912. DOI: 10.1038/srep25912
  177. 177. Lenzen S. Oxidative stress: the vulnerable beta-cell. Biochemical Society Transactions. 2008;36:343-347. DOI: 10.1042/bst0360343
  178. 178. Lenzen S. Chemistry and biology of reactive species with special reference to the antioxidative defence status in pancreatic β-cells. Biochimica et Biophysica Acta - General Subjects. 2017;1861:1929-1942. DOI: 10.1016/j.bbagen.2017.05.013
  179. 179. Lenzen S, Drinkgern J, Tiedge M. Low antioxidant enzyme gene expression in pancreatic islets compared with various other mouse tissues. Free Radical Biology & Medicine. 1996;20:463-466
  180. 180. Welsh N, Margulis B, Borg LA, Wiklund HJ, Saldeen J, Flodström M, et al. Differences in the expression of heat-shock proteins and antioxidant enzymes between human and rodent pancreatic islets: implications for the pathogenesis of insulin-dependent diabetes mellitus. Molecular medicine (Cambridge, Mass.). 1995;1:806-820
  181. 181. Ivarsson R, Quintens R, Dejonghe S, Tsukamoto K, in ’t Veld P, Renström E, et al. Redox control of exocytosis: regulatory role of NADPH, thioredoxin, and glutaredoxin. Diabetes. 2005;54:2132-2142
  182. 182. Reinbothe TM, Ivarsson R, Li D-Q , Niazi O, Jing X, Zhang E, et al. Glutaredoxin-1 Mediates NADPH-Dependent Stimulation of Calcium-Dependent Insulin Secretion. Molecular Endocrinology. 2009;23:893-900. DOI: 10.1210/me.2008-0306
  183. 183. Jezek P, Holendova B, Plecita-Hlavata L. Redox Signaling from Mitochondria: Signal Propagation and Its Targets. Biomolecules. 2020;10. DOI: 10.3390/biom10010093
  184. 184. Woo HA, Yim SH, Shin DH, Kang D, Yu DY, Rhee SG. Inactivation of peroxiredoxin I by phosphorylation allows localized H(2)O(2) accumulation for cell signaling. Cell. 2010;140:517-528. DOI: 10.1016/j.cell.2010.01.009

Written By

Petr Ježek, Blanka Holendová, Martin Jabůrek, Jan Tauber, Andrea Dlasková and Lydie Plecitá-Hlavatá

Submitted: 20 June 2020 Reviewed: 02 October 2020 Published: 18 June 2021