Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Aug 1.
Published in final edited form as: Free Radic Biol Med. 2014 May 28;0:383–399. doi: 10.1016/j.freeradbiomed.2014.05.016

REGULATION OF OBESITY AND INSULIN RESISTANCE BY NITRIC OXIDE

Brian E Sansbury 1,2, Bradford G Hill 1,2,3
PMCID: PMC4112002  NIHMSID: NIHMS600940  PMID: 24878261

Abstract

Obesity is a risk factor for developing type 2 diabetes and cardiovascular disease and has quickly become a world-wide pandemic with few tangible and safe treatment options. While it is generally accepted that the primary cause of obesity is energy imbalance, i.e., the calories consumed are greater than are utilized, understanding how caloric balance is regulated has proven a challenge. Many “distal” causes of obesity, such as the structural environment, occupation, and social influences, are exceedingly difficult to change or manipulate. Hence, molecular processes and pathways more proximal to the origins of obesity—those that directly regulate energy metabolism or caloric intake—appear to be more feasible targets for therapy. In particular, nitric oxide (NO) is emerging as a central regulator of energy metabolism and body composition. NO bioavailability is decreased in animal models of diet-induced obesity and in obese and insulin resistant patients, and increasing NO output has remarkable effects on obesity and insulin resistance. This review discusses the role of NO in regulating adiposity and insulin sensitivity and places its modes of action into context with the known causes and consequences of metabolic disease.

Keywords: nitric oxide, diabetes, insulin resistance, obesity, eNOS, mitochondria

1. The Obesity Epidemic

The recent increase in the prevalence of obesity is cause for alarm. The Centers for Disease Control (CDC) estimates that, from 1962 to 2010, obesity prevalence increased from 13% to 36%. As of 2008, approximately 1.5 billion adults aged 20 years or older were overweight, and 10% were obese [1]; more recent data from the United States indicate that >33% of adults and 17% of children are obese [2]. This has led to a dramatic increase in pre-diabetic states. For example, current estimates indicate that one-third of the population in the United States meets the criteria for pre-diabetes [3, 4], and, in addition to type 2 diabetes (T2D), obesity is closely associated with co-mordities such as cardiovascular disease, hypertension, atherosclerosis, stroke, and cancer [5]. Hence, the current high prevalence of obesity is likely to have a considerable impact on worldwide health. In addition, the economic burden of obesity is substantial and accounts for an estimated $147 billion per year in health care costs [6]. The problem has become so severe that, in 2013, the American Medical Association House of Delegates declared obesity a disease.

The principal cause of obesity is energy imbalance: the calories consumed are greater than that utilized by bodily processes, e.g., breathing, digestion, thermogenesis [7]. Indeed, the average consumption of calories in the US has increased by >200 kcal/d per person, which is partly attributable to the abundance of affordable, widely marketed, energy-dense foods [8-11]. Nevertheless, evidence suggests that the balance between calorie intake and energy expenditure is complex and regulated by many factors. Exposure to increasingly obesogenic environments has been suggested to promote not only overeating, but inactivity as well. For example, the human environment is fraught with both chemical and structural “obesogens.” These include but are not limited to: pollutants that promote adiposity and insulin resistance [12-21]; lack of structural features of the built environment that promote an active lifestyle, such as easy access to parks, sidewalks, and bike paths [22-24]; and the night/day cycles in the natural environment of the individual, which can be altered in those having certain occupations [25-27]. Moreover, the genetic makeup of individuals shows strong associations with the predisposition to become obese [28-30].

Many of these factors influence body composition in an indirect or distal manner, and thus could be considered “distal causes” of obesity (Figure 1). Intervening to address these distal causes is exceedingly difficult. For example, changing the structural environment would likely entail departing from particular types of communities or neighborhoods. Similarly, living under favorable day-night cycles is impossible for workers in some occupations, and changing genetic makeup is currently not an option. Even weight loss via caloric restriction faces difficulties, including an evolution-engendered guard against low fat mass [7, 31] and the propensity of the body to increase caloric efficiency during dieting [32, 33]. The intransigency of these problems has led to a search for causes more proximal to obesity, which may be tangible targets for anti-obesity therapies.

Figure 1. Distal and proximal causes of obesity.

Figure 1

Graphical illustration of the common causes of, or factors that contribute to, obesity: Influencing factors distal to the disease, such as policy as well as structural and chemical “obesogens” of the built and social (cultural) environment, may contribute to the prevalence of obesity. Funding for obesity research, dietary guidelines, physical education policies, and sidewalk standards are examples of potential influences related to Policy, which is most distal to the actual disease. The Built environment, which comprises places created or modified by people—i.e., where individuals work, their transportation systems, and life outside their homes—is another cause distal to obesity. The Social or cultural environment includes those family or cultural influences that affect behavioral activity, occupation (which may involve shift work), and social and media norms, all of which could affect eating habits and physical activity. Lastly, direct mechanisms that control hunger, satiety, energy expenditure, and nutrient absorption are Proximal causes of obesity. Commonly, these proximal causes are more tangible targets for anti-obesity/diabetes therapies compared with distal causes and are commonly regulated by nitric oxide.

2. Metabolic pathways known to regulate obesity

Understanding the mechanisms that promote adiposity and insulin resistance are critical to stem the growing tide of metabolic disease. In particular, the development of therapies for obesity and T2D requires a better understanding of the biochemical pathways that regulate metabolism and body composition. As a first principle, energy balance must be considered to understand how changes in body composition could occur. Any effective obesity treatment must decrease energy intake, increase energy expenditure or both. As discussed later, nitric oxide (NO) plays an important role in many of these proximal causes of obesity, which include:

  1. Hunger and satiety: The central nervous system regulates caloric intake and the feeling of satisfaction or fullness after a meal, i.e., satiety. This regulation is dependent on neural and endocrine inputs that can be divided into short- and long-term control systems. Release of cholecystokinin (CCK) in combination with neural signaling in response to gut distension are potent signals of satiety and trigger an end to feeding [34]. The adipose tissue-derived hormone, leptin, is crucial to integrate the melanocortin neuronal circuit of the hypothalamus with the energy stores of the body [34-36]. In addition to leptin, neuropeptide Y (NPY) directly affects feeding behavior, metabolism and body composition [37, 38], and corticotropin-releasing hormone, growth-hormone-releasing hormone, galanin and ghrelin, some of which are expressed in both the stomach and the brain, function in hunger and satiety signaling [39]. The neurotransmitters norepinephrine, dopamine and serotonin are also important in central energy balance [34, 36] and inhibiting their reuptake is the strategy employed by drugs such as sibutramine, which have proven effective but have side effects such as increased blood pressure and heart rate [40]. Other drugs that have been shown effective in decreasing energy intake by suppressing appetite are reviewed elsewhere [7, 41, 42].

  2. Nutrient absorption: Targeting nutrient absorption in the gut may be an effective obesity therapy. Signals from the gut released post-prandially are important not only in regulating food intake, but also in digestion and nutrient absorption. Ghrelin and CCK, as well as, peptide YY, glucagon-like peptides 1 and 2, gastric inhibitory peptide and corticotropin-releasing factor function to regulate both signaling and digestion [39, 43, 44]. Inhibition of gastric and pancreatic lipases via orlistat treatment decreases triglyceride hydrolysis and is able to inhibit absorption of ingested fat by ~30% and contributes to a caloric deficit of approximately 200 calories per day [45]. As with neurotransmitter reuptake inhibitors, orlistat promotes weight loss; however, side effects caused discontinuation of the medication in many patients [40].

  3. Energy expenditure: The largest contributor to obligatory energy expenditure is the basal metabolic rate (BMR), which is defined as the resting energy expenditure at thermoneutrality in the unfed state [46]. BMR includes cellular turnover, repair and basic functions (e.g., maintenance of ion gradients, transmembrane metabolite transfer), basal synthetic reactions (e.g., RNA, DNA protein synthesis) and mitochondrial proton leak; it also includes obligatory thermogenesis (e.g., digestion and absorption) [7, 46]. Mitochondria are central to the regulation of energy expenditure, and targeting their activity has been a prospect for obesity therapies for decades [47, 48], [49]. Therapies that mimic physiological, anti-obesogenic effects are likely to prove most effective. For example, overexpression of uncoupling protein 1 (UCP1), which increases substrate utilization and electron transport chain turnover, in white adipose tissue [50] or skeletal muscle [51] can prevent diet-induced obesity in mice, suggesting that uncoupling of oxidative phosphorylation in these two organs is sufficient to regulate body composition. Brown fat, which expresses relatively high levels of UCP1, is an exciting target for therapy. Despite the low amounts of brown fat in humans, as little as 50 g of brown fat has been estimated to be capable of utilizing up to 20% of basal caloric needs [52]; and mice with genetically reduced brown adipose tissue (BAT) mass are prone to obesity [53]. The recent discovery that adult humans maintain active depots of BAT [54, 55], in conjunction with the identification of UCP-2 and UCP-3 in the skeletal muscle and other tissues [56, 57], suggests that enhancement of mitochondrial activity may hold promise for combatting obesity.

    The finding that adipocytes in some white adipose tissue depots can be programmed to become similar to brown adipose tissue has further invigorated research into understanding the role of adipose tissue in systemic metabolism. A general idea is to increase mitochondria and UCP1 in white adipose tissue, which could promote a higher BMR. Several molecular targets have been identified, with PPAR-γ-coactivator-1α (PGC-1α) being of critical importance [58]. PGC-1α is a known regulator of energy metabolism and of mitochondrial biogenesis [59] and has been shown to induce many of the characteristic brown fat traits in white adipocytes in vitro [58]. Additionally, recent studies have identified secreted proteins that stimulate brown adipocyte thermogenesis and recruit brown (or beige) adipocytes to white adipose tissue [60] [61].

How can these pathways, as well as others that regulate factors proximal to the cause of obesity, be modulated to stem the tide of the obesity epidemic? Emerging evidence suggests that changes in vascular function could regulate metabolic homeostasis, and many studies show that the gaseous signaling molecule, nitric oxide (NO), may play a pivotal role in regulating systemic metabolism, body composition, and insulin sensitivity. In the sections that follow, we discuss NO, how it is regulated by dietary conditions and obesity, and its known effects on physiology, metabolism and adiposity.

3. Nitric oxide – endogenous formation and general modes of biological action

Nitric oxide and related nitrogen oxides have emerged as critical regulators of cell and tissue function [62]. The potency of NO was perhaps first realized following its inhalation by Sir Humphrey Davy, who nearly died from the self-experiment, and after which he vowed to “never design again…so rash an experiment” [63]. Nearly two centuries later, identification of the cardiovascular processes controlled by NO led to the Nobel Prize in Physiology or Medicine in 1998. Nevertheless, the pleiotropy of NO continues to unfold, and we are now beginning to appreciate the deeper aspects of its impact on metabolism.

Generation of NO

The most common route of NO production is through the action of the nitric oxide synthase (NOS) family of enzymes [62, 64]. NOS enzymes catalyze NADPH- and O2-dependent oxidation of L-arginine to L-citrulline, producing NO in the process. Such synthesis of NO depends on the availability of cofactors such as FAD, FMN, tetrahydrobiopterin (BH4), as well as the prosthetic group, heme [65].

The three isoforms of NOS generate NO at different rates [62]. NOS3 (commonly called endothelial NOS or eNOS) is expressed to highest relative abundance in the vascular endothelium, but has also been found in neurons, epithelial cells, and cardiomyocytes [66] as well as adipocytes [67, 68] and hepatocytes [69-71]. It produces relatively low quantities of NO, and its activity is controlled by Ca2+ and calmodulin, post-translational modifications [72, 73], and physical forces such as shear stress [74, 75]. NOS1 (frequently called neuronal NOS or nNOS) is also a Ca2+/calmodulin-dependent isoform that can be activated by agonists of the N-methyl-D-aspartate (NMDA) receptor [75]. It is expressed mostly in neurons, skeletal muscle, and epithelial cells. Lastly, NOS2 (also termed inducible NOS or iNOS), which has the highest capacity to generate NO, is expressed in multiple cell types in response to inflammatory stimuli [75, 76]. The results of some studies also suggest a mitochondria-localized NOS isoform [77, 78]; however, the specific contribution of this isoform remains unclear. In addition to post-translational modifications and substrate and cofactor availability, NOS activity is regulated by its localization within cells and by interactions with itself and other proteins [75].

In addition to direct production via NOS enzymes, NO can be produced endogenously from the more oxidized nitrogen oxide, nitrite. Reduction of nitrite to NO is increased under acidic and hypoxic conditions, with the reduction occurring enzymatically by heme proteins such as deoxyhemoglobin or deoxymyoglobin [79]. The therapeutic potential of dietary or pharmacological nitrite is supported by multiple studies describing improvements in reperfusion injury following myocardial infarction, in pulmonary hypertension, and injury after organ transplantation [80].

Biochemical properties of NO

NO is a free radical of rather limited biological reactivity. The endogenous half-life of NO has been found to be in the range of 2 ms to > 2 s and appears to depend on the availability of reactants [81]. NO primarily reacts with ferrous (Fe2+) iron and with other radical species; such reactions form the basis for nearly all of the biological effects of NO. The highest affinity interactions of NO are with metalloproteins such as soluble guanylate cyclase (sGC), cytochrome c oxidase, and hemoglobin; NO reacts also with non-heme iron. Reaction of NO with Fe2+ iron results in the formation of a coordinate bond, which is termed a nitrosyl adduct (i.e., nitrosylation). The presence of free radicals such as superoxide (O2.−) [82-85] changes the fate of NO: once reacted, it forms peroxynitrite [82, 84, 85] and can no longer bind to ferrous heme [86]. Peroxynitrite and reactive species derived from it (e.g., nitrogen dioxide) are important in inflammatory responses [85] and can modulate cell signaling [87-90], in part by promoting the oxidation or nitration of a broad range of biomolecules [62].

The NO molecule can also react directly with O2, which itself is a free radical possessing two unpaired electrons in different π* antibonding orbitals [91]. The reactions of NO with O2 commonly underlie the mechanisms by which S-nitrosation or S-oxidation of protein side chains occurs. In addition, NO can react directly with thiyl radicals, forming a nitroso covalent bond between NO and the thiol (termed S-nitrosation or S-nitrosylation). Cysteinyl thiols of glutathione and proteins are among the most recognized and studied targets of reactive nitrogen species (RNS). Reaction of RNS with thiols results in the formation of S-nitrosothiols, S-glutathiolated species, and oxidized cysteinyl residues [92]. Such modifications can lead to transient changes in enzyme activity, providing redox switches that can be modulated by addition or removal of the modifications [93].

General physiological roles of NO

NO has multiple biological actions and can regulate physiology acutely or lead to long-term changes in cell function. The pleiotropic roles of NO include the regulation of long-term synaptic transmission, learning, memory, platelet aggregation, leukocyte-endothelial interactions, immune function, and angiogenesis and arteriogenesis (for review, see [94]). However, NO is most well known as a potent regulator of blood flow and was originally termed endothelial-derived relaxing factor (EDRF). The story unfolded from Furchgott and Zawadzki’s initial discovery that endothelial cells control acetylcholine-induced relaxation of smooth muscle [95]. A few years later, NO was identified as the key endothelium-derived molecule promoting vasodilation: NO synthesized by NOS in the endothelium diffuses into the vessel wall where it activates sGC in vascular smooth muscle; this leads to a rise in cyclic GMP (cGMP) and elicits relaxation of the vessel [96-102]. However, it readily became apparent that different isoforms of NOS have different physiological functions. For example, NOS3 and NOS1 were found to have distinct roles in regulating microvascular tone [103]; NOS1 activity in the medulla and hypothalamus is important for systemic regulation of blood pressure [104-107]; and, the nitrergic nerves containing NOS1 are responsible for penile erection [99, 108]. Overall, NO derived from the integration of NOS3 and NOS1 activities play key roles in regulating systemic blood pressure and acutely regulating organ blood flow, whereas NOS2-derived NO species are most well recognized for their impact on pathogen killing and inflammatory processes [82].

Another key function of NO is the regulation of mitochondrial respiration. Acutely, NO inhibits respiration by binding and inhibiting cytochrome c oxidase. Modulation of respiration by NO is dependent on both mitochondrial activity and the O2 level [109, 110]. In addition, NO directly regulates the binding and release of oxygen with hemoglobin [111] and is able to increase blood flow at sites of very low oxygen concentrations [112]. Thus, a key function of NO is to modulate O2 gradients in cells and tissues by regulating hemoglobin action and by inhibiting O2 consumption in respiring mitochondria [81]. Chronic exposure to relatively high levels of NO results in mitochondrial biogenesis [113-115], which could reprogram a cell or tissue to have a higher metabolic capacity.

4. NO bioavailability is diminished in obese and diabetic states

NO bioavailability is decreased in animal models of obesity and diabetes [116, 117] and in obese and diabetic humans [118, 119]. Because its bioavailability is dependent upon the balance between its generation and degradation, diminished levels of NO in obese states may be due to decreased expression of NOS, impairments in NOS activity, or by the reaction of NO with reactive species (e.g., superoxide) (Figure 2). These are discussed in turn below.

Figure 2. Mechanisms for decreased endothelial-derived NO in obesity and diabetes.

Figure 2

Schematic of changes in NOS3 or NO: (A) Decreased NOS3 expression commonly occurs in obese and diabetic states. The mechanisms proposed for diminished expression include TNF-α-mediated destabilization of NOS3 mRNA, which may involve eEF1A1. High levels of NO may regulate NOS3 abundance through cGMP-mediated or via NF-κB-SNO feedback regulatory pathways. A small 27-nt RNA regulates NOS3 expression also, although it is not known whether this mechanism is invoked in obesity or diabetes. (B) Decreased NOS3 activity in obesity and diabetes is largely attributed to insulin resistance, which may be mediated by free fatty acid (FFA)-induced activation of TLR2, TLR4, and NF-κB. In addition, activation of PKCβII may diminish Akt signaling, which normally promotes phosphorylation of NOS3 on Ser1177. Phosphorylation at this site increases NO output by the enzyme. Hyperglycemia may also lead to increased O-GlcNAcylation of NOS3, which decreases Ser1177 phosphorylation and inhibits its activity. In addition, conditions leading to obesity promote upregulation of Cav-1, which is a negative regulator of NOS3, and ceramide accumulation disrupts the NOS3-Akt-HSP90 complex, diminishing activity of the enzyme. (C) NOS3 may also be uncoupled or NO quenched in obese and diabetic states. Diminished levels of substrates and cofactors, such as L-arginine or tetrahydrobiopterin (BH4), lead to uncoupling of the enzyme, which is commonly associated with the presence of NOS3 monomers rather than dimers and can produce superoxide instead of NO. Endogenous inhibitors of NOS3 such as ADMA are also increased in obese conditions and can promote NOS uncoupling. Elevated production of reactive oxygen species such as superoxide can quench NO and result in its oxidation to highly reactive peroxynitrite, which damages biomolecules and can oxidize BH4 to BH2.

NOS expression changes in obesity

A primary mechanism by which NO bioavailability could be decreased is via diminished expression of NOS enzymes (Figure 2A). In particular, lower NOS3 abundance is found in both adipose tissue and skeletal muscle of obese humans and rodents [120-124]. Factors associated with obesity and diabetes including shear stress, lysophosphatidylcholine, oxidized LDL, insulin and lack of exercise can also regulate NOS3 expression [125-129]. In multiple studies, tumor necrosis factor-α (TNFα), which is increased in obesity and implicated in the etiology of insulin resistance [130], was shown to downregulate NOS3 expression and abundance [123, 131-134] by decreasing the stability of NOS3 mRNA [135, 136], effectively shortening its half-life [137]. This destabilization of the NOS3 message was due, at least in part, to upregulation of elongation factor 1-α1 [138].

Acutely, TNF increases NOS3 activity [139], most likely via activation of the PI3KAkt [140] and sphingomyelinase/sphingosine-1-phosphate pathways [141, 142]. Such diametrically opposite acute versus chronic effects of TNF would appear to suggest the potential existence of negative feedback loops that sense high levels of NO, leading to downregulation of NOS3. Indeed, NO donors downregulate NOS3 expression both in vitro and in vivo, which may involve cGMP and/or S-nitros(yl)ation of NF-κB [143, 144]. A small, 27-nt RNA has also been shown to be an effective feedback regulator of NOS3 [145]. Whether such small RNAs or miRNAs that regulate NOS expression are induced with obesity is currently unclear.

Notable changes in the abundance of other NOS isoforms also occur in obesity. The NOS2 enzyme increases in abundance in pancreatic β-cells [146], aorta [147], skeletal muscle [148], liver [149, 150], and adipose tissue [151-153] of obese rodents. In adipose tissue, the majority of NOS2 is derived from infiltrating bone marrow-derived macrophages that display a proinflammatory phenotype [151-153]. Also, high levels of TNFα increase NOS2 in adipocytes, which appears to downregulate UCP-2 [154]; hence, this mechanism could contribute to decreases in white adipose tissue energy expenditure. In the ventromedial hypothalamus, which controls energy intake, diet-induced obesity was associated with lower numbers of NOS1-expressing cells [155]. In the aorta, however, NOS1 was increased in abundance in mice fed a high fat diet. The induction of NOS1 was demonstrated to be due to leptin stimulation [156] and may partially compensate for the loss of NOS3-mediated vasodilatory action that typically occurs in obese, insulin-resistant states.

Changes in NOS3 activity in obesity

Beyond changes in expression, the NO-producing activity of NOS3 is diminished in metabolic disease (Figure 2B). In addition to the required substrates, calcium, and cofactors, the activity of NOS3 is regulated by protein-protein interactions and by several post-translational modifications [94, 157, 158]. A high fat diet was shown to upregulate caveolin-1, a negative regulator of NOS3 [159, 160], in the aorta of obese rats [161]. Furthermore, ceramide (which is increased in obesity [162]) promotes disruption of the eNOS-Akt complex from HSP90 [163], which normally increases NOS3 activity by promoting displacement of caveolin-1 from NOS3 [164].

Conditions of obesity have profound effects on NOS3 phosphorylation. In particular, NOS3 phosphorylation at serine 1177 (S1177; S1176 in mice), which is critical for increasing NO output from the enzyme [165], is diminished in mice by nutrient excess [166-169] or high fat feeding [117, 122, 170, 171]; studies in obese rats [172-174] and pigs [175] show similar results. This NOS3 phosphorylation site can be regulated by Akt [176], which is activated by insulin [177]. Insulin stimulation of the Akt-NOS3 pathway could thus be important for regulating post-prandial blood flow and nutrient disposition to peripheral tissues. Indeed, insulin resistance in the endothelium is sufficient to diminish NO bioavailability and promote endothelial dysfunction [178], and impaired NOS3 phosphorylation due to insulin resistance was shown to be responsible for diminished glucose uptake in the skeletal muscle of mice subjected to nutrient excess [179].

The reasons for diminished phosphorylation of NOS3 under conditions of nutrient excess and obesity could be due to fatty acid (e.g., palmitate)-mediated induction of insulin resistance [117]. Elevated free fatty acids lower NO bioavailability in cultured cells [180], isolated arteries [181], animal models [182] and humans [183, 184] and this is likely due to decreased phosphoinositide 3-kinase (PI3K)-Akt-mediated activation of NOS3 [185]. Insulin resistance due to FFAs may be engendered by activation of Toll-like receptor 4 (TLR4) and NF-κB [170, 180] or Toll-like receptor 2 (TLR2) [186]. Other nutrient conditions inherent to diabetes may also be responsible for loss of S1177-NOS3 phosphorylation. For example, hyperglycemia causes O-linked N-acetylglucosamine (O-GlcNAc) modification of NOS3, which diminishes its activity [187]. Other mechanisms posited for diminished NOS3 phosphorylation in the context of obesity include a fatty acid-mediated, yet Akt-independent impairment of NOS3 phosphorylation [171], and PKCβII-mediated diminishment in Akt and NOS3 responsiveness to insulin [172, 174].

Interestingly, insulin-sensitizing therapeutic agents have robust effects on both NOS3 and NOS2. Several glitazone derivatives (also known as thiazolidinediones) used to treat type 2 diabetes increase NOS3 activity or expression in cultured endothelial cells [188-191]. Furthermore, the anti-diabetes drug metformin was shown to stimulate NOS3 activity in vitro and in vivo in an AMP kinase-dependent and HSP90-mediated manner [192]. Conversely, insulin-sensitizing PPARγ agonists robustly repress NOS2 expression in multiple cell and tissue types [193-198].

Uncoupling of NOS and quenching of NO in metabolic disease

The ability of NOS to produce NO is also dependent on its proper coupling, which is regulated by multiple cofactors, the ability of the NOS enzyme to remain in the dimerized form [199, 200],and post-translational modifications [94, 201-203] (Figure 2C). In particular, the cofactor BH4 is critical to NOS activity, and it has been termed a ‘redox sensor’ because elevations in reactive oxygen and nitrogen species can result in its depletion [204]. Furthermore, BH4 has been suggested to reflect the overall ‘health’ of the endothelium [205]. Obese and diabetic states in rodents and human cells are associated with decreased BH4 and elevated levels of its oxidized form, BH2 [205-209]. This is important because deficiency in BH4 or elevations in BH2 can uncouple NOS, which results in superoxide production from the enzyme and increases peroxynitrite generation [201]. Hence, deficiency of BH4 is thought to be a major regulator of vascular dysfunction that occurs during obesity and in diabetic states. Indeed, the ratio of BH4 to BH2 is critical in preventing glucose-induced NOS3 uncoupling [210] and replenishment of BH4 pools via either providing sepiapterin or by inducing pathways that increase biopterin synthesis or maintain BH4 in its reduced state have proven effective in multiple pathological scenarios (for review, see [201, 202, 211-213]). Uncoupling of NOS does not appear to be a factor unique to NOS3, however, as NOS1 was shown to be uncoupled in penile arteries of obese rats, leading to nitrergic dysfunction, which was corrected by increasing BH4 levels [214].

Peroxynitrite may be especially critical in promoting NOS uncoupling. 3-nitrotyrosine (3-NT) is a typical ‘footprint’ post-translational modification that helps identify the sites at which NOS3 uncoupling might have occurred, and it is worth noting that this modification is observed in abundance in the context of obesity and diabetes (e.g., [122, 206, 215, 216]). Patients with diabetes had diminished flow-mediated dilation of coronary arterioles and increased 3-NT protein adducts that colocalized with caveolae, demonstrating a dysfunction of the endothelium associated with elevated peroxynitrite production [217]. Interestingly, endothelial dysfunction in the diabetic patients was rescued by sepiapterin supplementation [217], inferring that peroxynitrite may disrupt NOS3 function not only by caveolar disruption, but by depleting BH4. This would be consistent with multiple studies showing that elevated levels of reactive species (in addition to peroxynitrite, such as superoxide produced from NADPH oxidase) promote NOS3 uncoupling [218-222]. Nevertheless, the specific contribution of peroxynitrite and other reactive species to endothelial function is still unclear, as other studies suggest that rather than uncoupling NOS3, superoxide from NADPH oxidase activates the enzyme; hence, inhibited NOS3 function perceived under conditions of oxidative stress could be due in part to the quenching of NO and not to uncoupling of the enzyme per se [223]. While this would be consistent with the near diffusion-limited reaction rate of NO with superoxide (which is reported to be as high as 1.9 × 1010 M−1 s−1) [224], the evidence for a deleterious role of uncoupled NOS should not be underestimated, and multiple other factors beyond BH4 depletion, such as asymmetric dimethyl arginine (ADMA), insufficient L-arginine levels or glutathio(ny)lation of the NOS3 enzyme, can promote NOS3 uncoupling and endothelial dysfunction [94, 225-228]. That levels of ADMA are positively correlated with insulin resistance and diabetes and that arginine supplementation overcomes this competitive inhibition [229] further suggests that losses in NOS3 coupling or activity are major contributors to the development of metabolic diseases associated with obesity.

Despite these findings, it is unclear whether obesity itself decreases NO availability. The fact that obesity in humans is associated with decreased blood flow in response to methacholine [230], bradykinin [231, 232], substance P and acetylcholine [232], shear stress [233], and insulin [234, 235] appears to suggest that the obese condition is somehow linked causally with diminished vascular NO bioavailability; a multitude of studies showing similar results lends support to this hypothesis [120, 236-247]. However, the question remains: Is loss of NO production somehow due to excess adiposity, or is its etiology derived from those conditions commonly associated with obesity? Interestingly, endothelial dysfunction was found to occur in morbidly obese humans only when insulin resistance was present [248]. And, severely obese humans, in the absence of insulin resistance, showed better flow-mediated dilation compared with normal and obese insulin-sensitive subjects [249]. Furthermore, capillary recruitment has been shown to be higher in overweight compared with lean individuals [250]. This suggests that the maintenance of a metabolically benign form of obesity is possible and that either insulin resistance or conditions directly linked with the insulin resistant phenotype (e.g., dyslipidemia, inflammation, hyperglycemia) are to blame for loss of NO bioavailability during obesity. Collectively, these findings bring forth multiple questions: What determines how metabolically benign versus harmful forms of obesity evolve?; How does NO affect obesity and insulin resistance?; and, What is the relevance of changes in NOS isoform abundance, some of which go in diametrically opposite directions (e.g., NOS3 vs. NOS2), in the development of metabolic disease?

5. Regulation of obesity and insulin resistance by NO

Both pharmacological studies and gain-of-function and loss-of-function studies have helped elucidate critical roles for NO in regulating obesity and insulin resistance. To date, supplementation with the NOS substrate, L-arginine, and inhibition of NOSs have been the most common pharmacological approaches used to determine how NO regulates body composition and insulin sensitivity. Genetic approaches, utilizing mice in which components integral to the synthesis of nitric oxide have been deleted or overexpressed, have further led to the development of a model by which NO regulates systemic metabolism. The model is complex and involves nearly all aspects thought to be important in regulating metabolic homeostasis.

Lessons from pharmacological interventions

Early pharmacological studies, using primarily L-arginine and NOS inhibitors, showed that NO is a potent regulator of both energy intake and expenditure. Interestingly, both L-arginine and inhibitors of NOS prevent obesity and insulin resistance, albeit by different mechanisms.

NO increases food intake

In rodents, L-arginine was shown to increase, and NOS inhibitors to decrease, food intake [155, 251-254]. These effects were due to NO activity in the brain, impinging on the leptin and serotonergic systems that regulate hunger. Leptin, given intracranially, was found to diminish diencephalic NOS activity and decrease food intake and body weight gain, and intracranial co-administration of L-arginine antagonized this effect [255]. Furthermore, intracerebroventricular injection of L-arginine, likely through stimulation of NOS activity, inhibited serotonin-induced anorexia caused by IL-1β [256]. Studies with NOS inhibitors have further solidified our understanding of the central effects of NO on hunger. Systemic administration of the NOS inhibitor, NG-nitro-L-arginine, reduced food intake in obese rats and increased serotonin metabolism in the cortex, diencephalon, and medulla pons, thereby implicating the central serotoninergic system in mediating the anorexic effect of NOS inhibitors [257]. Other NOS inhibitors, such as L-NAME, promote weight loss and diminish food intake in ob/ob and db/db mice [253] and obese rats [258] as well as reduce adiposity and improve insulin sensitivity in high fat-fed mouse models [259]. Interestingly, intracerebroventricular administration of NG-monomethyl-L-arginine (LNMMA) was shown also to regulate insulin secretion and peripheral insulin sensitivity [260], suggesting that centrally derived NO has effects that extend to distal nodes of systemic metabolism. It is also possible that this effect contributes to the hyperphagic effects of NO, as insulin is well known to regulate hunger and satiety [261-265]. Taken together with numerous other studies demonstrating a role for NO in the regulation of hunger [266-270], it would appear that NO produced in the brain antagonizes anorectic signals and stimulates food intake.

NO increases energy expenditure and regulates glucose and lipid metabolism

Ostensibly, this might imply that by promoting food intake, increased levels of NO, e.g., that elicited by supplementation with L-arginine, should increase adiposity and insulin resistance. However, human studies have repeatedly shown that L-arginine supplementation has favorable effects on body composition and insulin sensitivity [271-277]. Results from animal studies are in general agreement: L-arginine has been shown to have multimodal effects characterized by decreased fat mass, increased muscle mass, and improved insulin sensitivity. Despite promoting hyperphagia, L-arginine feeding reduced white adipose tissue mass, improved insulin sensitivity, and increased energy expenditure in mice [278]. In rats, not only has dietary L-arginine supplementation been shown to reduce fat mass, but it appears to increase skeletal muscle and brown fat mass and reduce serum concentrations of glucose, triglycerides, free fatty acids, homocysteine, dimethylarginines, and leptin as well [279, 280]. Similar salubrious systemic effects of L-arginine were demonstrated in pigs [281]. Overall, the collective data suggest that L-arginine, and by inference, NO, has the capacity to reduce fat mass by increasing mitochondrial biogenesis, regulating brown adipose tissue signaling, and increasing the expression of genes that promote oxidation of energy substrates [282].

Chronic treatment with sildenafil, which prevents the degradation of cGMP and is commonly prescribed to improve erectile function, improved insulin action and diminished obesity in high fat-fed mice [283]. Shorter durations of sildenafil treatment were shown to promote “browning” of white adipose tissue [284]. While there are no reports of regulation of obesity in humans by type 5-phosphodiesterase inhibitors such as sildenafil, they have been shown to increase mitochondrial biogenesis in human adipose tissue ex vivo [285], suggesting at least the potential to promote energy expenditure.

Other compounds that function in the NO pathway lend additional support to a role for NO in regulating insulin sensitivity. Restoration of NOS3 phosphorylation in endothelial-specific insulin receptor substrate 2 (IRS-2) knockout mice using beraprost (a stable prostaglandin analog) was sufficient to rescue capillary recruitment and to promote adequate insulin and glucose delivery to skeletal muscle [179]. Insulin, L-arginine, and sodium nitroprusside, by promoting S-nitro(sy)lation of key proteins, were shown to be particularly critical for regulating vascular endothelial insulin uptake and its transendothelial transport [286]. Hence, NO derived from NOS3 appears to be pivotal for regulating systemic glucose metabolism and insulin delivery to peripheral tissues.

As has been a theme with NO, its effects depend on its site of generation, concentration, and duration of application. Particularly interesting are its modes of action in the liver, skeletal muscle, and pancreas. Although chronic treatment with NOS inhibitors promote weight loss and insulin sensitivity in animal models [253, 258, 259], acute applications of inhibitors of NOS cause systemic insulin resistance [287]. This, in part, is mediated by actions in the liver, which can regulate systemic responses to insulin [288]. Administration of BH4, which is known to be oxidized to BH2 in the diabetic state [289-291] and plays an important role in regulating coupled NOS3 activity (vide supra), to STZ-induced diabetic mice lowered fasting blood glucose levels in an NOS3-dependent manner and improved glucose tolerance and insulin sensitivity in ob/ob mice [292]. This metabolic improvement was at least partially due to NOS3-mediated activation of AMPK in the liver, which suppressed hepatic gluconeogenesis [292]. Hence, the activity of NOS3 uncoupling in liver may be important for regulating systemic glucose metabolism.

Other studies also demonstrate an important role for liver NOS3 and nitrogen oxides. For example, intraportal administration of NOS inhibitors was shown to cause insulin resistance, which was rescued by intraportal delivery of the NO and superoxide donor, SIN-1 [293, 294]. Interestingly, when liver glutathione was first depleted by buthionine sulfoximine, the effects could not be rescued by sodium nitroprusside or SIN-1 [294]. These results suggest that nitrosated glutathione (GSNO) might be important in maximizing liver-mediated systemic responses to insulin. That intraportal delivery of glutathione methyl ester and SIN-1 enhances insulin sensitivity in rats would appear to support this hypothesis [295].

The NO-HISS connection?

This adds an additional question: How does NO (and its oxidation products) in the liver mediate systemic responses to insulin? The answer has been suggested to lie in a hormone called ‘hepatic insulin-sensitizing substance (HISS)’. This substance, for which there are only suggestive candidates (e.g., bone morphogenetic protein-9 [296]), has been suggested to account for 55% of the glucose disposal effect of insulin. Briefly, it is posited that post-prandial elevations in insulin result in release of a hormone, i.e., HISS, from the liver that acts on skeletal muscle to promote glucose uptake [297, 298]. Intriguingly, one study suggests that HISS, not insulin action, regulates the peripheral vasodilation generally attributed to insulin [299]. Atropine or hepatic surgical denervation inhibited the peripheral vascular actions of insulin, allegedly by blocking HISS release, and intraportal delivery of acetylcholine, which increases NOS activity, restored HISS release and insulin-mediated vasodilation [299]. This is consistent with original studies showing that insulin-mediated vasodilation is dependent on NO [300, 301] and that insulin-mediated skeletal muscle vasodilation contributes to insulin sensitivity in humans [302]; and, combined with other studies suggesting a role for NO in promoting release of HISS [303-305], this suggests that the putative hormone could be an NO-regulated, liver-produced, endocrine mediator of classical EDRF important for glucose disposal. What remains to be elucidated (in addition to the identity of HISS) is how this distally engendered mode of vasoregulation integrates physiologically (and pathologically) with the known local effects of insulin and NO in the vasculature [176-179].

Pancreatic effects of NO

Extremely important for maintaining metabolic homeostasis, the pancreas also utilizes NO to regulate its function. Briefly, the pancreas is comprised of two types of glands: (1) exocrine glands, which secrete the bicarbonate and digestive enzymes needed to neutralize the acidic gastric contents entering the small intestine and to complete digestion of food, respectively; and (2) endocrine glands, i.e., the islets of Langerhans, which contain several types of secretory cells, including α cells, β cells, δ cells, and F cells. Each of these cells secretes multiple proteins, such as insulin (β cells), glucagon (α cells), and somatostatin (δ cells). NO has been shown to affect both exocrine and endocrine functions of the pancreas. The effects of NO on the exocrine actions of the pancreas can be found in [306, 307].

With respect to insulin release, it appears that NO stimulates early, glucose-induced insulin release, while it is responsible for cytokine (e.g., IL-1β)-mediated inhibition of insulin secretion. This dual role of NO in regulating insulin secretion has been the subject of controversy (e.g., [308]), which might be based, in part, on the mechanistic complexity regulating pancreatic insulin secretion, and is likely compounded by the multiple actions of NO. The inhibitory actions of NO on insulin release appear to be due to NOS2-derived NO, which is implicated in the destruction of islet cells in type 1 diabetes [309, 310]. The mechanisms regulating the insulin-stimulating effects of NO [311-314] have been more difficult to elucidate. However, NO is suggested to stimulate islet cell insulin secretion by inducing calcium release from mitochondria [315], which may be due to NO-mediated inhibition of respiration and mitochondrial depolarization. Nevertheless, the stimulatory effects of NO on insulin secretion are relatively subtle [314], which might explain why other studies suggest that NO is not involved in the initiation of insulin secretion from pancreatic islets [316, 317].

Lessons from human studies and genetic interventions

Considerable data has accumulated suggesting an association between genetic polymorphisms in NOS isoforms and insulin resistance. Most notably, several studies have associated a T(−786)C variant of the NOS3 gene with insulin resistance [318-320]. Several other genetic variants in the NOS3 locus are associated with T2D [321], susceptibility for insulin resistance, hypertriglyceridemia, and low HDL [322], or worsened endothelial function in individuals prone to T2D [323]. Polymorphisms in the NOS2 gene have been associated with higher plasma glucose and elevated waist/hip ratios [324], and variants in the NOS2 gene promoter were found to be associated with T2D [325].

Genetic manipulation of NOS isoforms in mice have allowed for interrogation of the mechanisms by which NO regulates metabolic health and disease. In mice, it has been reported that deletion of NOS3 causes insulin resistance, hyperlipidemia, and hypertension [326]. While full gene deletion mimics human “metabolic syndrome,” even partial gene deletion of NOS3 results in exaggerated insulin resistance, glucose intolerance, and hypertension induced by high fat feeding [327, 328]. A study using NOS3/NOS1 double-knockout mice showed that deletion of NOS3 promotes insulin resistance in both skeletal muscle and liver, whereas NOS1 deletion impairs insulin sensitivity only in liver [329]. Mice lacking all isoforms of NOS, i.e., NOS3/NOS1/NOS2 triple knockout mice, demonstrate increased visceral obesity, hypertension, hypertriglyceridemia, and impaired glucose tolerance, and, it is interesting to note, that this is one of the few mouse strains to date to have spontaneous myocardial infarctions, apparently due to unstable coronary arteriosclerotic lesions [330]. That NOS is important to insulin sensitivity was further shown by studies in mice in which overexpression of dimethylarginine dimethylaminohydrolase—an enzyme that catalyzes the breakdown of the endogenous inhibitor of NOS, ADMA—increased insulin sensitivity [331].

Anti-obesogenic effects of NOS3

It appears that the metabolic phenotype elicited by insufficient levels of NOS3-derived NO relates directly to defects in intermediary metabolism in key peripheral tissues. Supporting evidence supplied by NOS3 KO mice include a markedly lower energy expenditure and decreases in mitochondrial content and fatty acid oxidation in muscle compared with WT mice [332]; and, as expected, NOS3 KO mice demonstrate an impaired ability to exercise [333]. Gain-of-function studies show a remarkable ability of NOS3 to regulate body composition and increase metabolism. Supplementation of NOS3 KO mice with nitrate, which can be serially reduced to nitrite and NO, reduced not only blood pressure, but visceral fat and triglycerides as well, thus reversing features of metabolic syndrome [334]. Furthermore, mice overexpressing NOS3 show an anti-obesogenic phenotype characterized by resistance to accumulation of white adipose tissue in response to a high fat diet, a higher metabolic rate, resistance to diet-induced hyperinsulinemia, and remarkably lower plasma levels of free fatty acids and triglycerides [335]. Similar results were obtained in an NOS3 phosphomimetic point mutant mouse model [336, 337]. Mutation of serine 1176 of NOS3 to an aspartic acid resulted in increased endothelial NO production as well as resistance to diet-induced weight gain and hyperinsulinemia; mutation of the residue to an alanine, which cannot be phosphorylated, resulted in insulin resistance and features of metabolic syndrome [157, 337].

How does NOS3 regulate metabolism and body composition? Several possibilities exist. Consistent changes in plasma lipids insinuate a central role of NOS3 in lipid oxidation or synthesis, e.g., NOS3 KO mice have elevated plasma levels triglycerides and free fatty acids compared with WT mice [326, 328], while NOS3 transgenic mice show diminished abundance of the lipids [122]. That these differences are due to modulation of fat oxidation capacity is suggested by studies showing a direct effect of NO on the capacity to oxidize fat. Not only do NOS3 KO mice show diminished fat oxidation capacity in skeletal muscle [332], but administration of a NOS inhibitor is sufficient to increase serum triglycerides and diminish hepatic fatty acid oxidation in rats [338], potentially by decreasing the activity of carnitine palmitoyl transferase [339]. Similar, NOS inhibitor-dependent decreases in fat oxidation capacity occur in heart [340]. In isolated hepatocytes, treatment with NO donors was shown to increase fatty acid oxidation in a cGMP-dependent manner by inhibiting acetyl CoA carboxylase (thereby decreasing production of malonyl CoA) and by stimulating carnitine palmitoyl transferase activity [341]. Interestingly, NO also inhibits fatty acid synthesis in hepatocytes [341], which is consistent with studies showing that NOS inhibitors [342] or genetic deletion of NOS3 increases lipid synthesis in liver [343]. In skeletal muscle, genetic deletion of NOS3 increases neolipogenic genes expression while downregulating genes involved in β-oxidation [332].

That genes involved in fatty acid oxidation are modulated by NO is consistent with data showing that overexpression of NOS3 increases expression of peroxisome proliferator activated receptor (PPAR)-α [122], which is well known to regulate lipid metabolism [344]. However, it is possible that NO regulates fat oxidation post-translationally as well. Recent studies show widespread S-nitrosation of multiple enzymes involved in intermediary metabolism. In particular, the liver enzyme, very long chain acyl-coA dehydrogenase (VLCAD) was shown to be nitrosated at Cys238, which increased the catalytic efficiency of the enzyme, and this modification was absent in NOS3 KO mice [345]. Collectively, these studies suggest that the powerful anti-obesity effects of NOS3-derived NO could be due to simultaneous increases and decreases in fat oxidation and synthesis, respectively.

NOS2 promotes insulin resistance

NO and RNS generated from the NOS2 enzyme could also regulate systemic metabolism. Most profound are its effects on insulin resistance. Although ablation of the NOS2 gene has no effect on weight gain, its absence was shown to improve glucose tolerance, normalize insulin sensitivity, and prevent derangements in PI3K/Akt signaling in high fat-fed mice [346]. Commonly, increases in NOS2 expression in skeletal muscle of obese mice are associated with increased S-nitrosation of the insulin receptor (IR), IR substrate-1 (IRS-1), and Akt, suggesting that nitrosative post-translational modifications of proteins in the insulin signaling pathway are responsible for NOS2-induced insulin resistance [347, 348]. The presence of NOS2 appears to decrease the abundance of IRS-1 via promoting its proteasomal degradation [148]. Interestingly, an acute bout of exercise was sufficient to downregulate NOS2 in high fat-fed rats as well as prevent S-nitrosation of proteins involved in insulin signaling, and administration of an inhibitor of NOS2 (L-N6-(1-iminoethyl)lysine; L-NIL) pheno-copied these effects [349]. Also, aspirin—which is one of the oldest known treatments for diabetes [350, 351] and improves blood glucose and insulin sensitivity in diabetic patients [352] and animal models of T2D [353]—inhibited NOS2-mediated S-nitrosation of IR, IRS-1, and Akt in skeletal muscle and improved insulin sensitivity [354].

In other tissues, NOS2 expression is also important for regulating insulin sensitivity. Selective overexpression of NOS2 in liver is sufficient to cause hepatic insulin resistance, hyperglycemia and hyperinsulinemia [355], and the use of an NOS2-specific inhibitor (L-NIL) reversed hyperglycemia, hyperinsulinemia, and insulin resistance in ob/ob mice [150]. In obesity, proinflammatory macrophages accumulating in adipose tissue are responsible for the majority of NOS2 expression [151-153] and may propagate the inflammatory signaling implicated in insulin resistance [130]. Importantly, the role of NOS2 in adipose tissue appears to differ remarkably from the canonical NO-cGMP pathway, as high fat diet-induced increases in proinflammatory cytokines and macrophage recruitment were attenuated by administration of sildenafil [356]. Interestingly, lack of NOS2 does not prevent age-induced insulin resistance [357], which suggests that not all insulin resistant states are created equal [358]. In line with this, mice lacking the NOS1 isoform show insulin resistance that appears to be due to a sympathetic, alpha-adrenergic mechanism [359].

6. Summary

Integration of results from these studies helps to form a model illustrating the role of NO in regulating obesity and insulin resistance (Figure 3). NO derived from NOS3 appears to have both anti-obesogenic and insulin-sensitizing properties. Its anti-obesogenic role stems from its ability to increase fat oxidation in peripheral tissues such as skeletal muscle, liver, and adipose tissue. As mentioned above, there is evidence that it also decreases lipid synthesis in liver. The impact of NOS3 on glucose metabolism and insulin sensitivity is underpinned by its capacity to increase transport of insulin and glucose to key peripheral tissues such as skeletal muscle and to regulate gluconeogenesis. Additionally, there may be implications for NOS3-mediated HISS release, which appears to enhance the vasodilatory properties of insulin. That NOS3 prevents hyperinsulinemia in two independent genetic gain-of-function studies [122, 337] further suggests that it could impact glucose metabolism by modulating insulin release. Other isoforms of NOS appear to promote deleterious changes in metabolism. In the brain, evidence suggests that NOS1-derived NO promotes hyperphagia. The NOS2 isoform promotes insulin resistance in both liver and skeletal muscle and is critical in inflammatory responses in multiple tissues, most notably, the adipose organ. Counter to NOS3, NOS2 appears to promote gluconeogenesis, and NOS2 has remarkable effects on cytokine-mediated insulin secretion. From these studies, it is apparent that NO is one of the most critical features regulating metabolism, body composition, and insulin sensitivity. Harnessing its beneficial metabolic actions is an exciting prospect for combatting metabolic disease.

Figure 3. Working model of the systemic effects of NO on obesity and metabolism.

Figure 3

Illustration of major organs and processes affected by NO and nitrogen oxides derived from NOS3, NOS1, and NOS2: The NOS3 isoform shows anti-obesogenic and insulin-sensitizing effects, which appears to be based in the ability of the enzyme to decrease lipid synthesis and promote fat oxidation in the liver and skeletal muscle. Additionally, NOS3 may be implicated in the secretion of hepatic insulin sensitizing substance (HISS), which might support insulin sensitivity in peripheral tissues such as skeletal muscle. NOS3 is important also for maximizing delivery of insulin and substrates to skeletal muscle, and this is likely critical in regulating insulin sensitivity and glucose tolerance. Through its actions in liver and pancreas, NOS3 may also suppress gluconeogenesis and prevent hyperinsulinemia, respectively. Additionally, NO increases the abundance of mitochondria and stimulates substrate oxidation capacity in adipose tissue, effectively promoting “browning” of white adipocytes. Conversely, other isoforms of NOS appear to have a more malevolent role in metabolism. NO derived from NOS1 promotes hyperphagia, and NOS2-derived nitrogen oxides can promote insulin resistance and inflammation in key peripheral tissues such as liver, skeletal muscle, and adipose tissue. In addition, NOS2 may affect glucose homeostasis by increasing glucose output from the liver and by impairing the endocrine activities of the pancreas.

HIGHLIGHTS.

  • Inline graphic Obesity is a major pandemic of the 21st century

  • Inline graphic Decreased eNOS (NOS3) activity and abundance are commonly observed in obesity

  • Inline graphic eNOS (NOS3)-derived NO has an anti-obesogenic effect

  • Inline graphic iNOS (NOS2) promotes insulin resistance

  • Inline graphic NO derived from iNOS (NOS1) appears to regulate appetite

ACKNOWLEDGEMENTS

The authors acknowledge support from NIH grants GM103492 and HL078825. The authors wish to thank Drs. Steven P. Jones and Aruni Bhatnagar for helpful suggestions. We are grateful to Thomas P. Gorton for his modified Vitruvian man illustration, which is shown in the Graphical Abstract.

NON-STANDARD ABBREVIATIONS AND ACRONYMS

3-NT

3-nitrotyrosine

ADMA

asymmetric dimethyl arginine

Akt

protein kinase B

AMP-activated protein kinase; BAT

brown adipose tissue

BH2

dihydrobiopterin

BH4

tetrahydrobiopterin

BMR

basal metabolic rate

Cav-1

caveolin-1

CCK

cholecystokinin

CDC

centers for disease control

cGMP

cyclic guanosine monophosphate

CoA

coenzyme A

EDRF

endothelial-derived relaxing factor

eEF1A1

elongation factor 1-α1

FAD

flavin adenine dinucleotide

FFA

free fatty acid

FMN

flavin mononucleotide

GSNO

nitrosated glutathione

HDL

high-density lipoprotein

HISS

hepatic insulin-sensitizing substance

HSP90

heat shock protein 90

IL-1β

interleukin-1 beta

IR

insulin receptor

IRS-1

insulin receptor substrate-1

Irs2

insulin receptor substrate 2

KO

knockout

L-NAME

L-NG-nitroarginine methyl ester

LNIL

L-N6-(1-iminoethyl)lysine

L-NMMA

NG-monomethyl-L-arginine

LDL

low-density lipoprotein

NADPH

nicotinamide adenine dinucleotide phosphate

NF-κB

nuclear factor kappa-light-chain-enhancer of activated B cells

NMDA

N-methyl-D-aspartate

NO

nitric oxide

NOS1

neuronal nitric oxide synthase

NOS2

nitric oxide synthase 2

NOS3

nitric oxide synthase 3

NPY

neuropeptide Y

O-GlcNAc

O-linked N-acetylglucosamine

O2.−

superoxide

PGC-1α

PPAR-γ-coactivator-1α

PI3K

phosphoinositide 3-kinase

PKCβII

protein kinase C isoform βII

PPAR

peroxisome proliferator activated receptor

RNS

reactive nitrogen species

sGC

soluble guanylate cyclase

SIN-1

3-morpholinosydnonimine

SNO

S-nitrosothiol

STZ

Streptozotocin

T2D

type 2 diabetes

TLR

Toll-like receptor

TNFα

tumor necrosis factor α

UCP

uncoupling protein

VLCAD

very long chain acyl-coA dehydrogenase

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

REFERENCES

  • [1].Ahima RS. Digging deeper into obesity. J Clin Invest. 2011;121:2076–2079. doi: 10.1172/JCI58719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [2].Ogden CL, Carroll MD, Kit BK, Flegal KM. Prevalence of obesity in the United States, 2009-2010. NCHS data brief. 2012:1–8. [PubMed] [Google Scholar]
  • [3].Ervin RB. Prevalence of metabolic syndrome among adults 20 years of age and over, by sex, age, race and ethnicity, and body mass index: United States, 2003-2006. Natl Health Stat Report. 2009:1–7. [PubMed] [Google Scholar]
  • [4].Roger VL, Go AS, Lloyd-Jones DM, Benjamin EJ, Berry JD, Borden WB, Bravata DM, Dai S, Ford ES, Fox CS, Fullerton HJ, Gillespie C, Hailpern SM, Heit JA, Howard VJ, Kissela BM, Kittner SJ, Lackland DT, Lichtman JH, Lisabeth LD, Makuc DM, Marcus GM, Marelli A, Matchar DB, Moy CS, Mozaffarian D, Mussolino ME, Nichol G, Paynter NP, Soliman EZ, Sorlie PD, Sotoodehnia N, Turan TN, Virani SS, Wong ND, Woo D, Turner MB. Heart Disease and Stroke Statistics--2012 Update: A Report From the American Heart Association. Circulation. 2011 doi: 10.1161/CIR.0b013e31823ac046. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [5].Calle EE, Thun MJ, Petrelli JM, Rodriguez C, Heath CW., Jr Body-mass index and mortality in a prospective cohort of U.S. adults. N Engl J Med. 1999;341:1097–1105. doi: 10.1056/NEJM199910073411501. [DOI] [PubMed] [Google Scholar]
  • [6].Zamosky L. The obesity epidemic. While America swallows $147 billion in obesity-related healthcare costs, physicians called on to confront the crisis. Medical economics. 2013;90:14–17. [PubMed] [Google Scholar]
  • [7].Tseng YH, Cypess AM, Kahn CR. Cellular bioenergetics as a target for obesity therapy. Nature Rev Drug Discov. 2010;9:465–482. doi: 10.1038/nrd3138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [8].Kant AK, Graubard BI. Eating out in America, 1987-2000: trends and nutritional correlates. Prev Med. 2004;38:243–249. doi: 10.1016/j.ypmed.2003.10.004. [DOI] [PubMed] [Google Scholar]
  • [9].Nielsen SJ, Popkin BM. Patterns and trends in food portion sizes, 1977-1998. JAMA. 2003;289:450–453. doi: 10.1001/jama.289.4.450. [DOI] [PubMed] [Google Scholar]
  • [10].Briefel RR, Johnson CL. Secular trends in dietary intake in the United States. Annu Rev Nutr. 2004;24:401–431. doi: 10.1146/annurev.nutr.23.011702.073349. [DOI] [PubMed] [Google Scholar]
  • [11].Hill JO, Peters JC. Environmental contributions to the obesity epidemic. Science. 1998;280:1371–1374. doi: 10.1126/science.280.5368.1371. [DOI] [PubMed] [Google Scholar]
  • [12].Jones OA, Maguire ML, Griffin JL. Environmental pollution and diabetes: a neglected association. Lancet. 2008;371:287–288. doi: 10.1016/S0140-6736(08)60147-6. [DOI] [PubMed] [Google Scholar]
  • [13].Smink A, Ribas-Fito N, Garcia R, Torrent M, Mendez MA, Grimalt JO, Sunyer J. Exposure to hexachlorobenzene during pregnancy increases the risk of overweight in children aged 6 years. Acta paediatrica. 2008;97:1465–1469. doi: 10.1111/j.1651-2227.2008.00937.x. [DOI] [PubMed] [Google Scholar]
  • [14].Sun Q, Yue P, Deiuliis JA, Lumeng CN, Kampfrath T, Mikolaj MB, Cai Y, Ostrowski MC, Lu B, Parthasarathy S, Brook RD, Moffatt-Bruce SD, Chen LC, Rajagopalan S. Ambient air pollution exaggerates adipose inflammation and insulin resistance in a mouse model of diet-induced obesity. Circulation. 2009;119:538–546. doi: 10.1161/CIRCULATIONAHA.108.799015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [15].Schell LM, Burnitz KK, Lathrop PW. Pollution and human biology. Ann Hum Biol. 2010;37:347–366. doi: 10.3109/03014461003705511. [DOI] [PubMed] [Google Scholar]
  • [16].Xu X, Yavar Z, Verdin M, Ying Z, Mihai G, Kampfrath T, Wang A, Zhong M, Lippmann M, Chen LC, Rajagopalan S, Sun Q. Effect of early particulate air pollution exposure on obesity in mice: role of p47phox. Arterioscler Thromb Vasc Biol. 2010;30:2518–2527. doi: 10.1161/ATVBAHA.110.215350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Yan YH, Chou CC, Lee CT, Liu JY, Cheng TJ. Enhanced insulin resistance in diet-induced obese rats exposed to fine particles by instillation. Inhal Toxicol. 2011;23:507–519. doi: 10.3109/08958378.2011.587472. [DOI] [PubMed] [Google Scholar]
  • [18].Bolton JL, Smith SH, Huff NC, Gilmour MI, Foster WM, Auten RL, Bilbo SD. Prenatal air pollution exposure induces neuroinflammation and predisposes offspring to weight gain in adulthood in a sex-specific manner. FASEB J. 2012;26:4743–4754. doi: 10.1096/fj.12-210989. [DOI] [PubMed] [Google Scholar]
  • [19].Rundle A, Hoepner L, Hassoun A, Oberfield S, Freyer G, Holmes D, Reyes M, Quinn J, Camann D, Perera F, Whyatt R. Association of childhood obesity with maternal exposure to ambient air polycyclic aromatic hydrocarbons during pregnancy. Am J Epidemiol. 2012;175:1163–1172. doi: 10.1093/aje/kwr455. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [20].Cupul-Uicab LA, Skjaerven R, Haug K, Melve KK, Engel SM, Longnecker MP. In utero exposure to maternal tobacco smoke and subsequent obesity, hypertension, and gestational diabetes among women in the MoBa cohort. Environ Health Persp. 2012;120:355–360. doi: 10.1289/ehp.1103789. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [21].Langer P, Ukropec J, Kocan A, Drobna B, Radikova Z, Huckova M, Imrich R, Gasperikova D, Klimes I, Trnovec T. Obesogenic and diabetogenic impact of high organochlorine levels (HCB, p,p’-DDE, PCBs) on inhabitants in the highly polluted Eastern Slovakia. Endocr Regul. 2014;48:17–24. doi: 10.4149/endo_2014_01_17. [DOI] [PubMed] [Google Scholar]
  • [22].Sallis JF, Floyd MF, Rodriguez DA, Saelens BE. Role of built environments in physical activity, obesity, and cardiovascular disease. Circulation. 2012;125:729–737. doi: 10.1161/CIRCULATIONAHA.110.969022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [23].Ding D, Sallis JF, Kerr J, Lee S, Rosenberg DE. Neighborhood environment and physical activity among youth a review. Am J Preventive Med. 2011;41:442–455. doi: 10.1016/j.amepre.2011.06.036. [DOI] [PubMed] [Google Scholar]
  • [24].Durand CP, Andalib M, Dunton GF, Wolch J, Pentz MA. A systematic review of built environment factors related to physical activity and obesity risk: implications for smart growth urban planning. Obesity Rev. 2011;12:e173–182. doi: 10.1111/j.1467-789X.2010.00826.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [25].Antunes LC, Levandovski R, Dantas G, Caumo W, Hidalgo MP. Obesity and shift work: chronobiological aspects. Nutr Res Rev. 2010;23:155–168. doi: 10.1017/S0954422410000016. [DOI] [PubMed] [Google Scholar]
  • [26].Garaulet M, Ordovas JM, Madrid JA. The chronobiology, etiology and pathophysiology of obesity. Int J Obesity. 2010;34:1667–1683. doi: 10.1038/ijo.2010.118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [27].Challet E. Circadian clocks, food intake, and metabolism. Prog Mol Biol Transl Sci. 2013;119:105–135. doi: 10.1016/B978-0-12-396971-2.00005-1. [DOI] [PubMed] [Google Scholar]
  • [28].Widen E, Lehto M, Kanninen T, Walston J, Shuldiner AR, Groop LC. Association of a polymorphism in the beta 3-adrenergic-receptor gene with features of the insulin resistance syndrome in Finns. N Engl J Med. 1995;333:348–351. doi: 10.1056/NEJM199508103330604. [DOI] [PubMed] [Google Scholar]
  • [29].Bouchard C, Tremblay A, Despres JP, Nadeau A, Lupien PJ, Theriault G, Dussault J, Moorjani S, Pinault S, Fournier G. The response to long-term overfeeding in identical twins. N Engl J Med. 1990;322:1477–1482. doi: 10.1056/NEJM199005243222101. [DOI] [PubMed] [Google Scholar]
  • [30].Bouchard C. The biological predisposition to obesity: beyond the thrifty genotype scenario. Int J Obes (Lond) 2007;31:1337–1339. doi: 10.1038/sj.ijo.0803610. [DOI] [PubMed] [Google Scholar]
  • [31].Wells JC. Thrift: a guide to thrifty genes, thrifty phenotypes and thrifty norms. Int J Obes. 2009;33:1331–1338. doi: 10.1038/ijo.2009.175. [DOI] [PubMed] [Google Scholar]
  • [32].Redman LM, Heilbronn LK, Martin CK, de Jonge L, Williamson DA, Delany JP, Ravussin E. Metabolic and behavioral compensations in response to caloric restriction: implications for the maintenance of weight loss. PLoS One. 2009;4:e4377. doi: 10.1371/journal.pone.0004377. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [33].Leibel RL, Rosenbaum M, Hirsch J. Changes in energy expenditure resulting from altered body weight. N Engl J Med. 1995;332:621–628. doi: 10.1056/NEJM199503093321001. [DOI] [PubMed] [Google Scholar]
  • [34].Spiegelman BM, Flier JS. Obesity and the regulation of energy balance. Cell. 2001;104:531–543. doi: 10.1016/s0092-8674(01)00240-9. [DOI] [PubMed] [Google Scholar]
  • [35].Ahima RS, Flier JS. Leptin. Annu Rev Physiol. 2000;62:413–437. doi: 10.1146/annurev.physiol.62.1.413. [DOI] [PubMed] [Google Scholar]
  • [36].Schwartz MW, Woods SC, Porte D, Jr., Seeley RJ, Baskin DG. Central nervous system control of food intake. Nature. 2000;404:661–671. doi: 10.1038/35007534. [DOI] [PubMed] [Google Scholar]
  • [37].Stanley BG, Kyrkouli SE, Lampert S, Leibowitz SF. Neuropeptide Y chronically injected into the hypothalamus: a powerful neurochemical inducer of hyperphagia and obesity. Peptides. 1986;7:1189–1192. doi: 10.1016/0196-9781(86)90149-x. [DOI] [PubMed] [Google Scholar]
  • [38].Billington CJ, Briggs JE, Harker S, Grace M, Levine AS. Neuropeptide Y in hypothalamic paraventricular nucleus: a center coordinating energy metabolism. Am J Physiol. 1994;266:R1765–1770. doi: 10.1152/ajpregu.1994.266.6.R1765. [DOI] [PubMed] [Google Scholar]
  • [39].Beck B, Pourie G. Ghrelin, neuropeptide Y, and other feeding-regulatory peptides active in the hippocampus: role in learning and memory. Nutr Rev. 2013;71:541–561. doi: 10.1111/nure.12045. [DOI] [PubMed] [Google Scholar]
  • [40].Bray GA, Tartaglia LA. Medicinal strategies in the treatment of obesity. Nature. 2000;404:672–677. doi: 10.1038/35007544. [DOI] [PubMed] [Google Scholar]
  • [41].Baretic M. Obesity drug therapy. Minerva Endocrinol. 2013;38:245–254. [PubMed] [Google Scholar]
  • [42].Rueda-Clausen CF, Padwal RS, Sharma AM. New pharmacological approaches for obesity management. Nat Rev Endocrinol. 2013;9:467–478. doi: 10.1038/nrendo.2013.113. [DOI] [PubMed] [Google Scholar]
  • [43].Suzuki K, Simpson KA, Minnion JS, Shillito JC, Bloom SR. The role of gut hormones and the hypothalamus in appetite regulation. Endocr J. 2010;57:359–372. doi: 10.1507/endocrj.k10e-077. [DOI] [PubMed] [Google Scholar]
  • [44].Delzenne NM, Cani PD, Neyrinck AM. Modulation of glucagon-like peptide 1 and energy metabolism by inulin and oligofructose: experimental data. J Nutr. 2007;137:2547S–2551S. doi: 10.1093/jn/137.11.2547S. [DOI] [PubMed] [Google Scholar]
  • [45].Guerciolini R. Mode of action of orlistat. Int J Obes Relat Metab Disord. 1997;21(Suppl 3):S12–23. [PubMed] [Google Scholar]
  • [46].Harper ME, Green K, Brand MD. The efficiency of cellular energy transduction and its implications for obesity. Annu Rev Nutr. 2008;28:13–33. doi: 10.1146/annurev.nutr.28.061807.155357. [DOI] [PubMed] [Google Scholar]
  • [47].Cutting WC, Rytand DA, Tainter ML. Relationship between Blood Cholesterol and Increased Metabolism from Dinitrophenol and Thyroid. J Clin Invest. 1934;13:547–552. doi: 10.1172/JCI100604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [48].Tainter ML, Cutting WC, Stockton AB. Use of Dinitrophenol in Nutritional Disorders : A Critical Survey of Clinical Results. Am J Public Health Nations Health. 1934;24:1045–1053. doi: 10.2105/ajph.24.10.1045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [49].Colman E. Dinitrophenol and obesity: an early twentieth-century regulatory dilemma. Regul Toxicol Pharmacol : RTP. 2007;48:115–117. doi: 10.1016/j.yrtph.2007.03.006. [DOI] [PubMed] [Google Scholar]
  • [50].Kopecky J, Rossmeisl M, Hodny Z, Syrovy I, Horakova M, Kolarova P. Reduction of dietary obesity in aP2-Ucp transgenic mice: mechanism and adipose tissue morphology. Am J Physiol. 1996;270:E776–786. doi: 10.1152/ajpendo.1996.270.5.E776. [DOI] [PubMed] [Google Scholar]
  • [51].Li B, Nolte LA, Ju JS, Han DH, Coleman T, Holloszy JO, Semenkovich CF. Skeletal muscle respiratory uncoupling prevents diet-induced obesity and insulin resistance in mice. Nat Med. 2000;6:1115–1120. doi: 10.1038/80450. [DOI] [PubMed] [Google Scholar]
  • [52].Rothwell NJ, Stock MJ. Luxuskonsumption, diet-induced thermogenesis and brown fat: the case in favour. Clin Sci (Lond) 1983;64:19–23. doi: 10.1042/cs0640019. [DOI] [PubMed] [Google Scholar]
  • [53].Lowell BB, Flier JS. Brown adipose tissue, beta 3-adrenergic receptors, and obesity. Annu Rev Med. 1997;48:307–316. doi: 10.1146/annurev.med.48.1.307. [DOI] [PubMed] [Google Scholar]
  • [54].Cypess AM, Lehman S, Williams G, Tal I, Rodman D, Goldfine AB, Kuo FC, Palmer EL, Tseng YH, Doria A, Kolodny GM, Kahn CR. Identification and importance of brown adipose tissue in adult humans. N Engl J Med. 2009;360:1509–1517. doi: 10.1056/NEJMoa0810780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [55].Virtanen KA, Lidell ME, Orava J, Heglind M, Westergren R, Niemi T, Taittonen M, Laine J, Savisto NJ, Enerback S, Nuutila P. Functional brown adipose tissue in healthy adults. N Engl J Med. 2009;360:1518–1525. doi: 10.1056/NEJMoa0808949. [DOI] [PubMed] [Google Scholar]
  • [56].Kozak LP, Harper ME. Mitochondrial uncoupling proteins in energy expenditure. Annu Rev Nutr. 2000;20:339–363. doi: 10.1146/annurev.nutr.20.1.339. [DOI] [PubMed] [Google Scholar]
  • [57].Vidal-Puig A, Solanes G, Grujic D, Flier JS, Lowell BB. UCP3: an uncoupling protein homologue expressed preferentially and abundantly in skeletal muscle and brown adipose tissue. Biochem Biophys Res Commun. 1997;235:79–82. doi: 10.1006/bbrc.1997.6740. [DOI] [PubMed] [Google Scholar]
  • [58].Puigserver P, Wu Z, Park CW, Graves R, Wright M, Spiegelman BM. A cold-inducible coactivator of nuclear receptors linked to adaptive thermogenesis. Cell. 1998;92:829–839. doi: 10.1016/s0092-8674(00)81410-5. [DOI] [PubMed] [Google Scholar]
  • [59].Handschin C, Spiegelman BM. Peroxisome proliferator-activated receptor gamma coactivator 1 coactivators, energy homeostasis, and metabolism. Endocr Rev. 2006;27:728–735. doi: 10.1210/er.2006-0037. [DOI] [PubMed] [Google Scholar]
  • [60].Serra D, Mera P, Malandrino MI, Mir JF, Herrero L. Mitochondrial fatty acid oxidation in obesity. Antioxid Redox Signal. 2013;19:269–284. doi: 10.1089/ars.2012.4875. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [61].Bostrom P, Wu J, Jedrychowski MP, Korde A, Ye L, Lo JC, Rasbach KA, Bostrom EA, Choi JH, Long JZ, Kajimura S, Zingaretti MC, Vind BF, Tu H, Cinti S, Hojlund K, Gygi SP, Spiegelman BM. A PGC1-alpha-dependent myokine that drives brown-fat-like development of white fat and thermogenesis. Nature. 2012;481:463–468. doi: 10.1038/nature10777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [62].Hill BG, Dranka BP, Bailey SM, Lancaster JR, Jr., Darley-Usmar VM. What part of NO don’t you understand? Some answers to the cardinal questions in nitric oxide biology. J Biol Chem. 2010;285:19699–19704. doi: 10.1074/jbc.R110.101618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [63].Sprigge JS. Sir Humphry Davy; his researches in respiratory physiology and his debt to Antoine Lavoisier. Anaesthesia. 2002;57:357–364. doi: 10.1046/j.1365-2044.2002.02414.x. [DOI] [PubMed] [Google Scholar]
  • [64].Alderton WK, Cooper CE, Knowles RG. Nitric oxide synthases: structure, function and inhibition. Biochem J. 2001;357:593–615. doi: 10.1042/0264-6021:3570593. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [65].Li H, Poulos TL. Structure-function studies on nitric oxide synthases. J Inorg Biochem. 2005;99:293–305. doi: 10.1016/j.jinorgbio.2004.10.016. [DOI] [PubMed] [Google Scholar]
  • [66].Dudzinski DM, Michel T. Life history of eNOS: partners and pathways. Cardiovasc Res. 2007;75:247–260. doi: 10.1016/j.cardiores.2007.03.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [67].Shen W, Hao J, Feng Z, Tian C, Chen W, Packer L, Shi X, Zang W, Liu J. Lipoamide or lipoic acid stimulates mitochondrial biogenesis in 3T3-L1 adipocytes via the endothelial NO synthase-cGMP-protein kinase G signalling pathway. Br J Pharmacol. 2011;162:1213–1224. doi: 10.1111/j.1476-5381.2010.01134.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [68].Tedesco L, Valerio A, Cervino C, Cardile A, Pagano C, Vettor R, Pasquali R, Carruba MO, Marsicano G, Lutz B, Pagotto U, Nisoli E. Cannabinoid type 1 receptor blockade promotes mitochondrial biogenesis through endothelial nitric oxide synthase expression in white adipocytes. Diabetes. 2008;57:2028–2036. doi: 10.2337/db07-1623. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [69].Mei Y, Thevananther S. Endothelial nitric oxide synthase is a key mediator of hepatocyte proliferation in response to partial hepatectomy in mice. Hepatology. 2011;54:1777–1789. doi: 10.1002/hep.24560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [70].Vazquez-Chantada M, Ariz U, Varela-Rey M, Embade N, Martinez-Lopez N, Fernandez-Ramos D, Gomez-Santos L, Lamas S, Lu SC, Martinez-Chantar ML, Mato JM. Evidence for LKB1/AMP-activated protein kinase/ endothelial nitric oxide synthase cascade regulated by hepatocyte growth factor, S-adenosylmethionine, and nitric oxide in hepatocyte proliferation. Hepatology. 2009;49:608–617. doi: 10.1002/hep.22660. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [71].Zimmermann H, Kurzen P, Klossner W, Renner EL, Marti U. Decreased constitutive hepatic nitric oxide synthase expression in secondary biliary fibrosis and its changes after Roux-en-Y choledocho-jejunostomy in the rat. J Hepatol. 1996;25:567–573. doi: 10.1016/s0168-8278(96)80218-2. [DOI] [PubMed] [Google Scholar]
  • [72].Oess S, Icking A, Fulton D, Govers R, Muller-Esterl W. Subcellular targeting and trafficking of nitric oxide synthases. Biochem J. 2006;396:401–409. doi: 10.1042/BJ20060321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [73].Sessa WC. eNOS at a glance. J Cell Sci. 2004;117:2427–2429. doi: 10.1242/jcs.01165. [DOI] [PubMed] [Google Scholar]
  • [74].Balligand JL, Feron O, Dessy C. eNOS activation by physical forces: from short-term regulation of contraction to chronic remodeling of cardiovascular tissues. Physiol Rev. 2009;89:481–534. doi: 10.1152/physrev.00042.2007. [DOI] [PubMed] [Google Scholar]
  • [75].Kone BC, Kuncewicz T, Zhang W, Yu ZY. Protein interactions with nitric oxide synthases: controlling the right time, the right place, and the right amount of nitric oxide. Am J Physiol Renal Physiol. 2003;285:F178–190. doi: 10.1152/ajprenal.00048.2003. [DOI] [PubMed] [Google Scholar]
  • [76].Kleinert H, Pautz A, Linker K, Schwarz PM. Regulation of the expression of inducible nitric oxide synthase. Eur J Pharmacol. 2004;500:255–266. doi: 10.1016/j.ejphar.2004.07.030. [DOI] [PubMed] [Google Scholar]
  • [77].Finocchietto PV, Franco MC, Holod S, Gonzalez AS, Converso DP, Arciuch VG, Serra MP, Poderoso JJ, Carreras MC. Mitochondrial nitric oxide synthase: a masterpiece of metabolic adaptation, cell growth, transformation, and death. Exp Biol Med (Maywood) 2009;234:1020–1028. doi: 10.3181/0902-MR-81. [DOI] [PubMed] [Google Scholar]
  • [78].Kato K, Giulivi C. Critical overview of mitochondrial nitric-oxide synthase. Front Biosci. 2006;11:2725–2738. doi: 10.2741/2002. [DOI] [PubMed] [Google Scholar]
  • [79].Vitturi DA, Patel RP. Current perspectives and challenges in understanding the role of nitrite as an integral player in nitric oxide biology and therapy. Free Radic Biol Med. 2011;51:805–812. doi: 10.1016/j.freeradbiomed.2011.05.037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [80].Lundberg JO, Gladwin MT, Ahluwalia A, Benjamin N, Bryan NS, Butler A, Cabrales P, Fago A, Feelisch M, Ford PC, Freeman BA, Frenneaux M, Friedman J, Kelm M, Kevil CG, Kim-Shapiro DB, Kozlov AV, Lancaster JR, Jr., Lefer DJ, McColl K, McCurry K, Patel RP, Petersson J, Rassaf T, Reutov VP, Richter-Addo GB, Schechter A, Shiva S, Tsuchiya K, van Faassen EE, Webb AJ, Zuckerbraun BS, Zweier JL, Weitzberg E. Nitrate and nitrite in biology, nutrition and therapeutics. Nat Chem Biol. 2009;5:865–869. doi: 10.1038/nchembio.260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [81].Thomas DD, Liu X, Kantrow SP, Lancaster JR., Jr The biological lifetime of nitric oxide: implications for the perivascular dynamics of NO and O2. Proc Natl Acad Sci U S A. 2001;98:355–360. doi: 10.1073/pnas.011379598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [82].Pacher P, Beckman JS, Liaudet L. Nitric oxide and peroxynitrite in health and disease. Physiol Rev. 2007;87:315–424. doi: 10.1152/physrev.00029.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [83].Trujillo M, Ferrer-Sueta G, Radi R. Peroxynitrite detoxification and its biologic implications. Antioxid Redox Signal. 2008;10:1607–1620. doi: 10.1089/ars.2008.2060. [DOI] [PubMed] [Google Scholar]
  • [84].Szabo C, Ischiropoulos H, Radi R. Peroxynitrite: biochemistry, pathophysiology and development of therapeutics. Nat Rev Drug Discov. 2007;6:662–680. doi: 10.1038/nrd2222. [DOI] [PubMed] [Google Scholar]
  • [85].Beckman JS. Understanding peroxynitrite biochemistry and its potential for treating human diseases. Arch Biochem Biophys. 2009;484:114–116. doi: 10.1016/j.abb.2009.03.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [86].McAndrew J, Patel RP, Jo H, Cornwell T, Lincoln T, Moellering D, White CR, Matalon S, Darley-Usmar V. The interplay of nitric oxide and peroxynitrite with signal transduction pathways: implications for disease. Semin Perinatol. 1997;21:351–366. doi: 10.1016/s0146-0005(97)80002-x. [DOI] [PubMed] [Google Scholar]
  • [87].Alvarez B, Radi R. Peroxynitrite reactivity with amino acids and proteins. Amino Acids. 2003;25:295–311. doi: 10.1007/s00726-003-0018-8. [DOI] [PubMed] [Google Scholar]
  • [88].Xu S, Ying J, Jiang B, Guo W, Adachi T, Sharov V, Lazar H, Menzoian J, Knyushko TV, Bigelow D, Schoneich C, Cohen RA. Detection of sequence-specific tyrosine nitration of manganese SOD and SERCA in cardiovascular disease and aging. Am J Physiol Heart Circ Physiol. 2006;290:H2220–2227. doi: 10.1152/ajpheart.01293.2005. [DOI] [PubMed] [Google Scholar]
  • [89].Koeck T, Stuehr DJ, Aulak KS. Mitochondria and regulated tyrosine nitration. Biochem Soc Trans. 2005;33:1399–1403. doi: 10.1042/BST0331399. [DOI] [PubMed] [Google Scholar]
  • [90].Aulak KS, Miyagi M, Yan L, West KA, Massillon D, Crabb JW, Stuehr DJ. Proteomic method identifies proteins nitrated in vivo during inflammatory challenge. Proc Natl Acad Sci U S A. 2001;98:12056–12061. doi: 10.1073/pnas.221269198. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [91].Halliwell B, Gutteridge JMC. Free radicals in biology and medicine. Oxford University Press; Oxford: 1998. [Google Scholar]
  • [92].Hogg N. The biochemistry and physiology of S-nitrosothiols. Annu Rev Pharmacol Toxicol. 2002;42:585–600. doi: 10.1146/annurev.pharmtox.42.092501.104328. [DOI] [PubMed] [Google Scholar]
  • [93].Hill BG, Bhatnagar A. Role of glutathiolation in preservation, restoration and regulation of protein function. IUBMB life. 2007;59:21–26. doi: 10.1080/15216540701196944. [DOI] [PubMed] [Google Scholar]
  • [94].Forstermann U, Sessa WC. Nitric oxide synthases: regulation and function. Eur Heart J. 2012;33:829–837. 837a–837d. doi: 10.1093/eurheartj/ehr304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [95].Furchgott RF, Zawadzki JV. The obligatory role of endothelial cells in the relaxation of arterial smooth muscle by acetylcholine. Nature. 1980;288:373–376. doi: 10.1038/288373a0. [DOI] [PubMed] [Google Scholar]
  • [96].Rapoport RM, Draznin MB, Murad F. Endothelium-dependent relaxation in rat aorta may be mediated through cyclic GMP-dependent protein phosphorylation. Nature. 1983;306:174–176. doi: 10.1038/306174a0. [DOI] [PubMed] [Google Scholar]
  • [97].Rapoport RM, Murad F. Agonist-induced endothelium-dependent relaxation in rat thoracic aorta may be mediated through cGMP. Circ Res. 1983;52:352–357. doi: 10.1161/01.res.52.3.352. [DOI] [PubMed] [Google Scholar]
  • [98].Forstermann U, Mulsch A, Bohme E, Busse R. Stimulation of soluble guanylate cyclase by an acetylcholine-induced endothelium-derived factor from rabbit and canine arteries. Circ Res. 1986;58:531–538. doi: 10.1161/01.res.58.4.531. [DOI] [PubMed] [Google Scholar]
  • [99].Rajfer J, Aronson WJ, Bush PA, Dorey FJ, Ignarro LJ. Nitric oxide as a mediator of relaxation of the corpus cavernosum in response to nonadrenergic, noncholinergic neurotransmission. N Engl J Med. 1992;326:90–94. doi: 10.1056/NEJM199201093260203. [DOI] [PubMed] [Google Scholar]
  • [100].Moncada S, Palmer RM, Gryglewski RJ. Mechanism of action of some inhibitors of endothelium-derived relaxing factor. Proc Natl Acad Sci U S A. 1986;83:9164–9168. doi: 10.1073/pnas.83.23.9164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [101].Gryglewski RJ, Moncada S, Palmer RM. Bioassay of prostacyclin and endothelium-derived relaxing factor (EDRF) from porcine aortic endothelial cells. Br J Pharmacol. 1986;87:685–694. doi: 10.1111/j.1476-5381.1986.tb14586.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [102].Palmer RM, Ferrige AG, Moncada S. Nitric oxide release accounts for the biological activity of endothelium-derived relaxing factor. Nature. 1987;327:524–526. doi: 10.1038/327524a0. [DOI] [PubMed] [Google Scholar]
  • [103].Melikian N, Seddon MD, Casadei B, Chowienczyk PJ, Shah AM. Neuronal nitric oxide synthase and human vascular regulation. Trends Cardiovas Med. 2009;19:256–262. doi: 10.1016/j.tcm.2010.02.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [104].Togashi H, Sakuma I, Yoshioka M, Kobayashi T, Yasuda H, Kitabatake A, Saito H, Gross SS, Levi R. A central nervous system action of nitric oxide in blood pressure regulation. J Pharmacol Exp Ther. 1992;262:343–347. [PubMed] [Google Scholar]
  • [105].Sakuma I, Togashi H, Yoshioka M, Saito H, Yanagida M, Tamura M, Kobayashi T, Yasuda H, Gross SS, Levi R. NG-methyl-L-arginine, an inhibitor of L-arginine-derived nitric oxide synthesis, stimulates renal sympathetic nerve activity in vivo. A role for nitric oxide in the central regulation of sympathetic tone? Circ Res. 1992;70:607–611. doi: 10.1161/01.res.70.3.607. [DOI] [PubMed] [Google Scholar]
  • [106].el Karib AO, Sheng J, Betz AL, Malvin RL. The central effects of a nitric oxide synthase inhibitor (N omega-nitro-L-arginine) on blood pressure and plasma renin. Clin Exp Hypertens. 1993;15:819–832. doi: 10.3109/10641969309041644. [DOI] [PubMed] [Google Scholar]
  • [107].Toda N, Ayajiki K, Okamura T. Control of systemic and pulmonary blood pressure by nitric oxide formed through neuronal nitric oxide synthase. J Hypertens. 2009;27:1929–1940. doi: 10.1097/HJH.0b013e32832e8ddf. [DOI] [PubMed] [Google Scholar]
  • [108].Kim N, Azadzoi KM, Goldstein I, Saenz de Tejada I. A nitric oxide-like factor mediates nonadrenergic-noncholinergic neurogenic relaxation of penile corpus cavernosum smooth muscle. J Clin Invest. 1991;88:112–118. doi: 10.1172/JCI115266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [109].Shiva S, Oh JY, Landar AL, Ulasova E, Venkatraman A, Bailey SM, Darley-Usmar VM. Nitroxia: the pathological consequence of dysfunction in the nitric oxide-cytochrome c oxidase signaling pathway. Free Radic Biol Med. 2005;38:297–306. doi: 10.1016/j.freeradbiomed.2004.10.037. [DOI] [PubMed] [Google Scholar]
  • [110].Cooper CE, Giulivi C. Nitric oxide regulation of mitochondrial oxygen consumption II: Molecular mechanism and tissue physiology. Am J Physiol Cell Physiol. 2007;292:C1993–2003. doi: 10.1152/ajpcell.00310.2006. [DOI] [PubMed] [Google Scholar]
  • [111].Wolzt M, MacAllister RJ, Davis D, Feelisch M, Moncada S, Vallance P, Hobbs AJ. Biochemical characterization of S-nitrosohemoglobin. Mechanisms underlying synthesis, no release, and biological activity. J Biol Chem. 1999;274:28983–28990. doi: 10.1074/jbc.274.41.28983. [DOI] [PubMed] [Google Scholar]
  • [112].Patel RP, Hogg N, Spencer NY, Kalyanaraman B, Matalon S, Darley-Usmar VM. Biochemical characterization of human S-nitrosohemoglobin. Effects on oxygen binding and transnitrosation. J Biol Chem. 1999;274:15487–15492. doi: 10.1074/jbc.274.22.15487. [DOI] [PubMed] [Google Scholar]
  • [113].Kelly DP, Scarpulla RC. Transcriptional regulatory circuits controlling mitochondrial biogenesis and function. Genes Dev. 2004;18:357–368. doi: 10.1101/gad.1177604. [DOI] [PubMed] [Google Scholar]
  • [114].Nisoli E, Clementi E, Paolucci C, Cozzi V, Tonello C, Sciorati C, Bracale R, Valerio A, Francolini M, Moncada S, Carruba MO. Mitochondrial biogenesis in mammals: the role of endogenous nitric oxide. Science. 2003;299:896–899. doi: 10.1126/science.1079368. [DOI] [PubMed] [Google Scholar]
  • [115].Nisoli E, Falcone S, Tonello C, Cozzi V, Palomba L, Fiorani M, Pisconti A, Brunelli S, Cardile A, Francolini M, Cantoni O, Carruba MO, Moncada S, Clementi E. Mitochondrial biogenesis by NO yields functionally active mitochondria in mammals. Proc Natl Acad Sci U S A. 2004;101:16507–16512. doi: 10.1073/pnas.0405432101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [116].Bender SB, Herrick EK, Lott ND, Klabunde RE. Diet-induced obesity and diabetes reduce coronary responses to nitric oxide due to reduced bioavailability in isolated mouse hearts. Diabetes Obes Metab. 2007;9:688–696. doi: 10.1111/j.1463-1326.2006.00650.x. [DOI] [PubMed] [Google Scholar]
  • [117].Kim F, Pham M, Maloney E, Rizzo NO, Morton GJ, Wisse BE, Kirk EA, Chait A, Schwartz MW. Vascular inflammation, insulin resistance, and reduced nitric oxide production precede the onset of peripheral insulin resistance. Arterioscler Thromb Vasc Biol. 2008;28:1982–1988. doi: 10.1161/ATVBAHA.108.169722. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [118].Higashi Y, Sasaki S, Nakagawa K, Matsuura H, Chayama K, Oshima T. Effect of obesity on endothelium-dependent, nitric oxide-mediated vasodilation in normotensive individuals and patients with essential hypertension. Am J Hypertens. 2001;14:1038–1045. doi: 10.1016/s0895-7061(01)02191-4. [DOI] [PubMed] [Google Scholar]
  • [119].Gruber HJ, Mayer C, Mangge H, Fauler G, Grandits N, Wilders-Truschnig M. Obesity reduces the bioavailability of nitric oxide in juveniles. Int J Obes. 2008;32:826–831. doi: 10.1038/sj.ijo.0803795. [DOI] [PubMed] [Google Scholar]
  • [120].Georgescu A, Popov D, Constantin A, Nemecz M, Alexandru N, Cochior D, Tudor A. Dysfunction of human subcutaneous fat arterioles in obesity alone or obesity associated with Type 2 diabetes. Clin Sci. 2011;120:463–472. doi: 10.1042/CS20100355. [DOI] [PubMed] [Google Scholar]
  • [121].Perez-Matute P, Neville MJ, Tan GD, Frayn KN, Karpe F. Transcriptional control of human adipose tissue blood flow. Obesity. 2009;17:681–688. doi: 10.1038/oby.2008.606. [DOI] [PubMed] [Google Scholar]
  • [122].Sansbury BE, Cummins TD, Tang Y, Hellmann J, Holden CR, Harbeson MA, Chen Y, Patel RP, Spite M, Bhatnagar A, Hill BG. Overexpression of Endothelial Nitric Oxide Synthase Prevents Diet-Induced Obesity and Regulates Adipocyte Phenotype. Circ Res. 2012;111:1176–1189. doi: 10.1161/CIRCRESAHA.112.266395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [123].Valerio A, Cardile A, Cozzi V, Bracale R, Tedesco L, Pisconti A, Palomba L, Cantoni O, Clementi E, Moncada S, Carruba MO, Nisoli E. TNF-alpha downregulates eNOS expression and mitochondrial biogenesis in fat and muscle of obese rodents. J Clin Invest. 2006;116:2791–2798. doi: 10.1172/JCI28570. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [124].Kraus RM, Houmard JA, Kraus WE, Tanner CJ, Pierce JR, Choi MD, Hickner RC. Obesity, insulin resistance, and skeletal muscle nitric oxide synthase. J Appl Physiol (1985) 2012;113:758–765. doi: 10.1152/japplphysiol.01018.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [125].Awolesi MA, Sessa WC, Sumpio BE. Cyclic strain upregulates nitric oxide synthase in cultured bovine aortic endothelial cells. J Clin Invest. 1995;96:1449–1454. doi: 10.1172/JCI118181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [126].Zembowicz A, Tang JL, Wu KK. Transcriptional induction of endothelial nitric oxide synthase type III by lysophosphatidylcholine. J Biol Chem. 1995;270:17006–17010. doi: 10.1074/jbc.270.28.17006. [DOI] [PubMed] [Google Scholar]
  • [127].Hirata K, Miki N, Kuroda Y, Sakoda T, Kawashima S, Yokoyama M. Low concentration of oxidized low-density lipoprotein and lysophosphatidylcholine upregulate constitutive nitric oxide synthase mRNA expression in bovine aortic endothelial cells. Circ Res. 1995;76:958–962. doi: 10.1161/01.res.76.6.958. [DOI] [PubMed] [Google Scholar]
  • [128].Zeng G, Nystrom FH, Ravichandran LV, Cong LN, Kirby M, Mostowski H, Quon MJ. Roles for insulin receptor, PI3-kinase, and Akt in insulin-signaling pathways related to production of nitric oxide in human vascular endothelial cells. Circulation. 2000;101:1539–1545. doi: 10.1161/01.cir.101.13.1539. [DOI] [PubMed] [Google Scholar]
  • [129].Hambrecht R, Wolf A, Gielen S, Linke A, Hofer J, Erbs S, Schoene N, Schuler G. Effect of exercise on coronary endothelial function in patients with coronary artery disease. N Engl J Med. 2000;342:454–460. doi: 10.1056/NEJM200002173420702. [DOI] [PubMed] [Google Scholar]
  • [130].Hotamisligil GS, Shargill NS, Spiegelman BM. Adipose expression of tumor necrosis factor-alpha: direct role in obesity-linked insulin resistance. Science. 1993;259:87–91. doi: 10.1126/science.7678183. [DOI] [PubMed] [Google Scholar]
  • [131].Michel T, Lamas S. Molecular cloning of constitutive endothelial nitric oxide synthase: evidence for a family of related genes. J Cardiovasc Pharmacol. 1992;20(Suppl 12):S45–49. doi: 10.1097/00005344-199204002-00014. [DOI] [PubMed] [Google Scholar]
  • [132].Lai PF, Mohamed F, Monge JC, Stewart DJ. Downregulation of eNOS mRNA expression by TNFalpha: identification and functional characterization of RNA-protein interactions in the 3′UTR. Cardiovasc Res. 2003;59:160–168. doi: 10.1016/s0008-6363(03)00296-7. [DOI] [PubMed] [Google Scholar]
  • [133].Neumann P, Gertzberg N, Johnson A. TNF-alpha induces a decrease in eNOS promoter activity. Am J Physiol Lung Cell Mol Physiol. 2004;286:L452–459. doi: 10.1152/ajplung.00378.2002. [DOI] [PubMed] [Google Scholar]
  • [134].Anderson HD, Rahmutula D, Gardner DG. Tumor necrosis factor-alpha inhibits endothelial nitric-oxide synthase gene promoter activity in bovine aortic endothelial cells. J Biol Chem. 2004;279:963–969. doi: 10.1074/jbc.M309552200. [DOI] [PubMed] [Google Scholar]
  • [135].Alonso J, Sanchez de Miguel L, Monton M, Casado S, Lopez-Farre A. Endothelial cytosolic proteins bind to the 3′ untranslated region of endothelial nitric oxide synthase mRNA: regulation by tumor necrosis factor alpha. Mol Cell Biol. 1997;17:5719–5726. doi: 10.1128/mcb.17.10.5719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [136].Sanchez de Miguel L, Alonso J, Gonzalez-Fernandez F, de la Osada J, Monton M, Rodriguez-Feo JA, Guerra JI, Arriero MM, Rico L, Casado S, Lopez-Farre A. Evidence that an endothelial cytosolic protein binds to the 3′-untranslated region of endothelial nitric oxide synthase mRNA. J Vasc Res. 1999;36:201–208. doi: 10.1159/000025643. [DOI] [PubMed] [Google Scholar]
  • [137].Yoshizumi M, Perrella MA, Burnett JC, Jr., Lee ME. Tumor necrosis factor downregulates an endothelial nitric oxide synthase mRNA by shortening its half-life. Circ Res. 1993;73:205–209. doi: 10.1161/01.res.73.1.205. [DOI] [PubMed] [Google Scholar]
  • [138].Yan G, You B, Chen SP, Liao JK, Sun J. Tumor necrosis factor-alpha downregulates endothelial nitric oxide synthase mRNA stability via translation elongation factor 1-alpha 1. Circ Res. 2008;103:591–597. doi: 10.1161/CIRCRESAHA.108.173963. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [139].Bulotta S, Barsacchi R, Rotiroti D, Borgese N, Clementi E. Activation of the endothelial nitric-oxide synthase by tumor necrosis factor-alpha. A novel feedback mechanism regulating cell death. J Biol Chem. 2001;276:6529–6536. doi: 10.1074/jbc.M006535200. [DOI] [PubMed] [Google Scholar]
  • [140].Kawanaka H, Jones MK, Szabo IL, Baatar D, Pai R, Tsugawa K, Sugimachi K, Sarfeh IJ, Tarnawski AS. Activation of eNOS in rat portal hypertensive gastric mucosa is mediated by TNF-alpha via the PI 3-kinase-Akt signaling pathway. Hepatology. 2002;35:393–402. doi: 10.1053/jhep.2002.30958. [DOI] [PubMed] [Google Scholar]
  • [141].Barsacchi R, Perrotta C, Bulotta S, Moncada S, Borgese N, Clementi E. Activation of endothelial nitric-oxide synthase by tumor necrosis factor-alpha: a novel pathway involving sequential activation of neutral sphingomyelinase, phosphatidylinositol-3′ kinase, and Akt. Mol Pharmacol. 2003;63:886–895. doi: 10.1124/mol.63.4.886. [DOI] [PubMed] [Google Scholar]
  • [142].De Palma C, Meacci E, Perrotta C, Bruni P, Clementi E. Endothelial nitric oxide synthase activation by tumor necrosis factor alpha through neutral sphingomyelinase 2, sphingosine kinase 1, and sphingosine 1 phosphate receptors: a novel pathway relevant to the pathophysiology of endothelium. Arterioscler Thromb Vasc Biol. 2006;26:99–105. doi: 10.1161/01.ATV.0000194074.59584.42. [DOI] [PubMed] [Google Scholar]
  • [143].Vaziri ND, Wang XQ. cGMP-mediated negative-feedback regulation of endothelial nitric oxide synthase expression by nitric oxide. Hypertension. 1999;34:1237–1241. doi: 10.1161/01.hyp.34.6.1237. [DOI] [PubMed] [Google Scholar]
  • [144].Grumbach IM, Chen W, Mertens SA, Harrison DG. A negative feedback mechanism involving nitric oxide and nuclear factor kappa-B modulates endothelial nitric oxide synthase transcription. J Mol Cell Cardiol. 2005;39:595–603. doi: 10.1016/j.yjmcc.2005.06.012. [DOI] [PubMed] [Google Scholar]
  • [145].Zhang MX, Ou H, Shen YH, Wang J, Wang J, Coselli J, Wang XL. Regulation of endothelial nitric oxide synthase by small RNA. Proc Natl Acad Sci U S A. 2005;102:16967–16972. doi: 10.1073/pnas.0503853102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [146].Shimabukuro M, Ohneda M, Lee Y, Unger RH. Role of nitric oxide in obesity-induced beta cell disease. J Clin Invest. 1997;100:290–295. doi: 10.1172/JCI119534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [147].Noronha BT, Li JM, Wheatcroft SB, Shah AM, Kearney MT. Inducible nitric oxide synthase has divergent effects on vascular and metabolic function in obesity. Diabetes. 2005;54:1082–1089. doi: 10.2337/diabetes.54.4.1082. [DOI] [PubMed] [Google Scholar]
  • [148].Sugita H, Fujimoto M, Yasukawa T, Shimizu N, Sugita M, Yasuhara S, Martyn JA, Kaneki M. Inducible nitric-oxide synthase and NO donor induce insulin receptor substrate-1 degradation in skeletal muscle cells. J Biol Chem. 2005;280:14203–14211. doi: 10.1074/jbc.M411226200. [DOI] [PubMed] [Google Scholar]
  • [149].Mantena SK, Vaughn DP, Andringa KK, Eccleston HB, King AL, Abrams GA, Doeller JE, Kraus DW, Darley-Usmar VM, Bailey SM. High fat diet induces dysregulation of hepatic oxygen gradients and mitochondrial function in vivo. Biochem J. 2009;417:183–193. doi: 10.1042/BJ20080868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [150].Fujimoto M, Shimizu N, Kunii K, Martyn JA, Ueki K, Kaneki M. A role for iNOS in fasting hyperglycemia and impaired insulin signaling in the liver of obese diabetic mice. Diabetes. 2005;54:1340–1348. doi: 10.2337/diabetes.54.5.1340. [DOI] [PubMed] [Google Scholar]
  • [151].Weisberg SP, McCann D, Desai M, Rosenbaum M, Leibel RL, Ferrante AW., Jr Obesity is associated with macrophage accumulation in adipose tissue. J Clin Invest. 2003;112:1796–1808. doi: 10.1172/JCI19246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [152].Lumeng CN, Bodzin JL, Saltiel AR. Obesity induces a phenotypic switch in adipose tissue macrophage polarization. J Clin Invest. 2007;117:175–184. doi: 10.1172/JCI29881. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [153].Lumeng CN, Deyoung SM, Bodzin JL, Saltiel AR. Increased inflammatory properties of adipose tissue macrophages recruited during diet-induced obesity. Diabetes. 2007;56:16–23. doi: 10.2337/db06-1076. [DOI] [PubMed] [Google Scholar]
  • [154].Merial C, Bouloumie A, Trocheris V, Lafontan M, Galitzky J. Nitric oxide-dependent downregulation of adipocyte UCP-2 expression by tumor necrosis factor-alpha. Am J Physiol Cell Physiol. 2000;279:C1100–1106. doi: 10.1152/ajpcell.2000.279.4.C1100. [DOI] [PubMed] [Google Scholar]
  • [155].Sadler CJ, Wilding JP. Reduced ventromedial hypothalamic neuronal nitric oxide synthase and increased sensitivity to NOS inhibition in dietary obese rats: further evidence of a role for nitric oxide in the regulation of energy balance. Brain Res. 2004;1016:222–228. doi: 10.1016/j.brainres.2004.05.007. [DOI] [PubMed] [Google Scholar]
  • [156].Benkhoff S, Loot AE, Pierson I, Sturza A, Kohlstedt K, Fleming I, Shimokawa H, Grisk O, Brandes RP, Schroder K. Leptin potentiates endothelium-dependent relaxation by inducing endothelial expression of neuronal NO synthase. Arterioscler Thromb Vasc Biol. 2012;32:1605–1612. doi: 10.1161/ATVBAHA.112.251140. [DOI] [PubMed] [Google Scholar]
  • [157].Huang PL. eNOS, metabolic syndrome and cardiovascular disease. Trends Endocrin Met. 2009;20:295–302. doi: 10.1016/j.tem.2009.03.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [158].Rafikov R, Fonseca FV, Kumar S, Pardo D, Darragh C, Elms S, Fulton D, Black SM. eNOS activation and NO function: structural motifs responsible for the posttranslational control of endothelial nitric oxide synthase activity. J Endocrinol. 2011;210:271–284. doi: 10.1530/JOE-11-0083. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [159].Ju H, Zou R, Venema VJ, Venema RC. Direct interaction of endothelial nitric-oxide synthase and caveolin-1 inhibits synthase activity. J Biol Chem. 1997;272:18522–18525. doi: 10.1074/jbc.272.30.18522. [DOI] [PubMed] [Google Scholar]
  • [160].Michel JB, Feron O, Sacks D, Michel T. Reciprocal regulation of endothelial nitric-oxide synthase by Ca2+-calmodulin and caveolin. J Biol Chem. 1997;272:15583–15586. doi: 10.1074/jbc.272.25.15583. [DOI] [PubMed] [Google Scholar]
  • [161].Yang N, Ying C, Xu M, Zuo X, Ye X, Liu L, Nara Y, Sun X. High-fat diet up-regulates caveolin-1 expression in aorta of diet-induced obese but not in diet-resistant rats. Cardiovascular research. 2007;76:167–174. doi: 10.1016/j.cardiores.2007.05.028. [DOI] [PubMed] [Google Scholar]
  • [162].Bikman BT, Summers SA. Ceramides as modulators of cellular and whole-body metabolism. J Clin Invest. 2011;121:4222–4230. doi: 10.1172/JCI57144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [163].Zhang QJ, Holland WL, Wilson L, Tanner JM, Kearns D, Cahoon JM, Pettey D, Losee J, Duncan B, Gale D, Kowalski CA, Deeter N, Nichols A, Deesing M, Arrant C, Ruan T, Boehme C, McCamey DR, Rou J, Ambal K, Narra KK, Summers SA, Abel ED, Symons JD. Ceramide mediates vascular dysfunction in diet-induced obesity by PP2A-mediated dephosphorylation of the eNOS-Akt complex. Diabetes. 2012;61:1848–1859. doi: 10.2337/db11-1399. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [164].Gratton JP, Fontana J, O’Connor DS, Garcia-Cardena G, McCabe TJ, Sessa WC. Reconstitution of an endothelial nitric-oxide synthase (eNOS), hsp90, and caveolin-1 complex in vitro. Evidence that hsp90 facilitates calmodulin stimulated displacement of eNOS from caveolin-1. J Biol Chem. 2000;275:22268–22272. doi: 10.1074/jbc.M001644200. [DOI] [PubMed] [Google Scholar]
  • [165].McCabe TJ, Fulton D, Roman LJ, Sessa WC. Enhanced electron flux and reduced calmodulin dissociation may explain “calcium-independent” eNOS activation by phosphorylation. J Biol Chem. 2000;275:6123–6128. doi: 10.1074/jbc.275.9.6123. [DOI] [PubMed] [Google Scholar]
  • [166].Elrod JW, Duranski MR, Langston W, Greer JJ, Tao L, Dugas TR, Kevil CG, Champion HC, Lefer DJ. eNOS gene therapy exacerbates hepatic ischemia-reperfusion injury in diabetes: a role for eNOS uncoupling. Circ Res. 2006;99:78–85. doi: 10.1161/01.RES.0000231306.03510.77. [DOI] [PubMed] [Google Scholar]
  • [167].Zhong JC, Yu XY, Huang Y, Yung LM, Lau CW, Lin SG. Apelin modulates aortic vascular tone via endothelial nitric oxide synthase phosphorylation pathway in diabetic mice. Cardiovasc Res. 2007;74:388–395. doi: 10.1016/j.cardiores.2007.02.002. [DOI] [PubMed] [Google Scholar]
  • [168].Li Q, Atochin D, Kashiwagi S, Earle J, Wang A, Mandeville E, Hayakawa K, d’Uscio LV, Lo EH, Katusic Z, Sessa W, Huang PL. Deficient eNOS phosphorylation is a mechanism for diabetic vascular dysfunction contributing to increased stroke size. Stroke. 2013;44:3183–3188. doi: 10.1161/STROKEAHA.113.002073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [169].Taguchi K, Kobayashi T, Matsumoto T, Kamata K. Dysfunction of endothelium-dependent relaxation to insulin via PKC-mediated GRK2/Akt activation in aortas of ob/ob mice. Am J Physiol Heart Circ Physiol. 2011;301:H571–583. doi: 10.1152/ajpheart.01189.2010. [DOI] [PubMed] [Google Scholar]
  • [170].Kim F, Pham M, Luttrell I, Bannerman DD, Tupper J, Thaler J, Hawn TR, Raines EW, Schwartz MW. Toll-like receptor-4 mediates vascular inflammation and insulin resistance in diet-induced obesity. Circ Res. 2007;100:1589–1596. doi: 10.1161/CIRCRESAHA.106.142851. [DOI] [PubMed] [Google Scholar]
  • [171].Symons JD, McMillin SL, Riehle C, Tanner J, Palionyte M, Hillas E, Jones D, Cooksey RC, Birnbaum MJ, McClain DA, Zhang QJ, Gale D, Wilson LJ, Abel ED. Contribution of insulin and Akt1 signaling to endothelial nitric oxide synthase in the regulation of endothelial function and blood pressure. Circ Res. 2009;104:1085–1094. doi: 10.1161/CIRCRESAHA.108.189316. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [172].Naruse K, Rask-Madsen C, Takahara N, Ha SW, Suzuma K, Way KJ, Jacobs JR, Clermont AC, Ueki K, Ohshiro Y, Zhang J, Goldfine AB, King GL. Activation of vascular protein kinase C-beta inhibits Akt-dependent endothelial nitric oxide synthase function in obesity-associated insulin resistance. Diabetes. 2006;55:691–698. doi: 10.2337/diabetes.55.03.06.db05-0771. [DOI] [PubMed] [Google Scholar]
  • [173].Zecchin HG, Priviero FB, Souza CT, Zecchin KG, Prada PO, Carvalheira JB, Velloso LA, Antunes E, Saad MJ. Defective insulin and acetylcholine induction of endothelial cell-nitric oxide synthase through insulin receptor substrate/Akt signaling pathway in aorta of obese rats. Diabetes. 2007;56:1014–1024. doi: 10.2337/db05-1147. [DOI] [PubMed] [Google Scholar]
  • [174].Park K, Li Q, Rask-Madsen C, Mima A, Mizutani K, Winnay J, Maeda Y, D’Aquino K, White MF, Feener EP, King GL. Serine phosphorylation sites on IRS2 activated by angiotensin II and protein kinase C to induce selective insulin resistance in endothelial cells. Mol Cell Biol. 2013;33:3227–3241. doi: 10.1128/MCB.00506-13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [175].Low Wang CC, Lu L, Leitner JW, Sarraf M, Gianani R, Draznin B, Greyson CR, Reusch JE, Schwartz GG. Arterial insulin resistance in Yucatan micropigs with diet-induced obesity and metabolic syndrome. J Diabetes Complicat. 2013;27:307–315. doi: 10.1016/j.jdiacomp.2013.02.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [176].Dimmeler S, Fleming I, Fisslthaler B, Hermann C, Busse R, Zeiher AM. Activation of nitric oxide synthase in endothelial cells by Akt-dependent phosphorylation. Nature. 1999;399:601–605. doi: 10.1038/21224. [DOI] [PubMed] [Google Scholar]
  • [177].Hermann C, Assmus B, Urbich C, Zeiher AM, Dimmeler S. Insulin-mediated stimulation of protein kinase Akt: A potent survival signaling cascade for endothelial cells. Arterioscler Thromb Vasc Biol. 2000;20:402–409. doi: 10.1161/01.atv.20.2.402. [DOI] [PubMed] [Google Scholar]
  • [178].Duncan ER, Crossey PA, Walker S, Anilkumar N, Poston L, Douglas G, Ezzat VA, Wheatcroft SB, Shah AM, Kearney MT. Effect of endothelium-specific insulin resistance on endothelial function in vivo. Diabetes. 2008;57:3307–3314. doi: 10.2337/db07-1111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [179].Kubota T, Kubota N, Kumagai H, Yamaguchi S, Kozono H, Takahashi T, Inoue M, Itoh S, Takamoto I, Sasako T, Kumagai K, Kawai T, Hashimoto S, Kobayashi T, Sato M, Tokuyama K, Nishimura S, Tsunoda M, Ide T, Murakami K, Yamazaki T, Ezaki O, Kawamura K, Masuda H, Moroi M, Sugi K, Oike Y, Shimokawa H, Yanagihara N, Tsutsui M, Terauchi Y, Tobe K, Nagai R, Kamata K, Inoue K, Kodama T, Ueki K, Kadowaki T. Impaired insulin signaling in endothelial cells reduces insulin-induced glucose uptake by skeletal muscle. Cell Metab. 2011;13:294–307. doi: 10.1016/j.cmet.2011.01.018. [DOI] [PubMed] [Google Scholar]
  • [180].Kim F, Tysseling KA, Rice J, Pham M, Haji L, Gallis BM, Baas AS, Paramsothy P, Giachelli CM, Corson MA, Raines EW. Free fatty acid impairment of nitric oxide production in endothelial cells is mediated by IKKbeta. Arterioscler Thromb Vasc Biol. 2005;25:989–994. doi: 10.1161/01.ATV.0000160549.60980.a8. [DOI] [PubMed] [Google Scholar]
  • [181].Edirisinghe I, McCormick Hallam K, Kappagoda CT. Effect of fatty acids on endothelium-dependent relaxation in the rabbit aorta. Clin Sci. 2006;111:145–151. doi: 10.1042/CS20060001. [DOI] [PubMed] [Google Scholar]
  • [182].Du X, Edelstein D, Obici S, Higham N, Zou MH, Brownlee M. Insulin resistance reduces arterial prostacyclin synthase and eNOS activities by increasing endothelial fatty acid oxidation. J Clin Invest. 2006;116:1071–1080. doi: 10.1172/JCI23354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [183].Steinberg HO, Paradisi G, Hook G, Crowder K, Cronin J, Baron AD. Free fatty acid elevation impairs insulin-mediated vasodilation and nitric oxide production. Diabetes. 2000;49:1231–1238. doi: 10.2337/diabetes.49.7.1231. [DOI] [PubMed] [Google Scholar]
  • [184].Steinberg HO, Tarshoby M, Monestel R, Hook G, Cronin J, Johnson A, Bayazeed B, Baron AD. Elevated circulating free fatty acid levels impair endothelium-dependent vasodilation. J Clin Invest. 1997;100:1230–1239. doi: 10.1172/JCI119636. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [185].Yu Q, Gao F, Ma XL. Insulin says NO to cardiovascular disease. Cardiovascular research. 2011;89:516–524. doi: 10.1093/cvr/cvq349. [DOI] [PubMed] [Google Scholar]
  • [186].Jang HJ, Kim HS, Hwang DH, Quon MJ, Kim JA. Toll-like receptor 2 mediates high-fat diet-induced impairment of vasodilator actions of insulin. Am J Physiol Endocrinol Metab. 2013;304:E1077–1088. doi: 10.1152/ajpendo.00578.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [187].Du XL, Edelstein D, Dimmeler S, Ju Q, Sui C, Brownlee M. Hyperglycemia inhibits endothelial nitric oxide synthase activity by posttranslational modification at the Akt site. J Clin Invest. 2001;108:1341–1348. doi: 10.1172/JCI11235. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [188].Calnek DS, Mazzella L, Roser S, Roman J, Hart CM. Peroxisome proliferator-activated receptor gamma ligands increase release of nitric oxide from endothelial cells. Arterioscler Thromb Vasc Biol. 2003;23:52–57. doi: 10.1161/01.atv.0000044461.01844.c9. [DOI] [PubMed] [Google Scholar]
  • [189].Cho DH, Choi YJ, Jo SA, Jo I. Nitric oxide production and regulation of endothelial nitric-oxide synthase phosphorylation by prolonged treatment with troglitazone: evidence for involvement of peroxisome proliferator-activated receptor (PPAR) gamma-dependent and PPARgamma-independent signaling pathways. J Biol Chem. 2004;279:2499–2506. doi: 10.1074/jbc.M309451200. [DOI] [PubMed] [Google Scholar]
  • [190].Polikandriotis JA, Mazzella LJ, Rupnow HL, Hart CM. Peroxisome proliferator-activated receptor gamma ligands stimulate endothelial nitric oxide production through distinct peroxisome proliferator-activated receptor gamma-dependent mechanisms. Arterioscler Thromb Vasc Biol. 2005;25:1810–1816. doi: 10.1161/01.ATV.0000177805.65864.d4. [DOI] [PubMed] [Google Scholar]
  • [191].Goya K, Sumitani S, Otsuki M, Xu X, Yamamoto H, Kurebayashi S, Saito H, Kouhara H, Kasayama S. The thiazolidinedione drug troglitazone up-regulates nitric oxide synthase expression in vascular endothelial cells. J Diabetes Complicat. 2006;20:336–342. doi: 10.1016/j.jdiacomp.2005.08.003. [DOI] [PubMed] [Google Scholar]
  • [192].Davis BJ, Xie Z, Viollet B, Zou MH. Activation of the AMP-activated kinase by antidiabetes drug metformin stimulates nitric oxide synthesis in vivo by promoting the association of heat shock protein 90 and endothelial nitric oxide synthase. Diabetes. 2006;55:496–505. doi: 10.2337/diabetes.55.02.06.db05-1064. [DOI] [PubMed] [Google Scholar]
  • [193].Li M, Pascual G, Glass CK. Peroxisome proliferator-activated receptor gamma-dependent repression of the inducible nitric oxide synthase gene. Mol Cell Biol. 2000;20:4699–4707. doi: 10.1128/mcb.20.13.4699-4707.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [194].Maggi LB, Jr., Sadeghi H, Weigand C, Scarim AL, Heitmeier MR, Corbett JA. Anti-inflammatory actions of 15-deoxy-delta 12,14-prostaglandin J2 and troglitazone: evidence for heat shock-dependent and -independent inhibition of cytokine-induced inducible nitric oxide synthase expression. Diabetes. 2000;49:346–355. doi: 10.2337/diabetes.49.3.346. [DOI] [PubMed] [Google Scholar]
  • [195].Heneka MT, Klockgether T, Feinstein DL. Peroxisome proliferator-activated receptor-gamma ligands reduce neuronal inducible nitric oxide synthase expression and cell death in vivo. J Neurosci. 2000;20:6862–6867. doi: 10.1523/JNEUROSCI.20-18-06862.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [196].Konturek PC, Brzozowski T, Kania J, Konturek SJ, Kwiecien S, Pajdo R, Hahn EG. Pioglitazone, a specific ligand of peroxisome proliferator-activated receptor-gamma, accelerates gastric ulcer healing in rat. Eur J Pharmacol. 2003;472:213–220. doi: 10.1016/s0014-2999(03)01932-0. [DOI] [PubMed] [Google Scholar]
  • [197].Mendez M, LaPointe MC. PPARgamma inhibition of cyclooxygenase-2, PGE2 synthase, and inducible nitric oxide synthase in cardiac myocytes. Hypertension. 2003;42:844–850. doi: 10.1161/01.HYP.0000085332.69777.D1. [DOI] [PubMed] [Google Scholar]
  • [198].Crosby MB, Svenson JL, Zhang J, Nicol CJ, Gonzalez FJ, Gilkeson GS. Peroxisome proliferation-activated receptor (PPAR)gamma is not necessary for synthetic PPARgamma agonist inhibition of inducible nitric-oxide synthase and nitric oxide. J Pharmacol Exp Ther. 2005;312:69–76. doi: 10.1124/jpet.104.074005. [DOI] [PubMed] [Google Scholar]
  • [199].Rodriguez-Crespo I, Gerber NC, Ortiz de Montellano PR. Endothelial nitric-oxide synthase. Expression in Escherichia coli, spectroscopic characterization, and role of tetrahydrobiopterin in dimer formation. J Biol Chem. 1996;271:11462–11467. doi: 10.1074/jbc.271.19.11462. [DOI] [PubMed] [Google Scholar]
  • [200].Rodriguez-Crespo I, Ortiz de Montellano PR. Human endothelial nitric oxide synthase: expression in Escherichia coli, coexpression with calmodulin, and characterization. Arch Biochem Biophys. 1996;336:151–156. doi: 10.1006/abbi.1996.0543. [DOI] [PubMed] [Google Scholar]
  • [201].Alp NJ, Channon KM. Regulation of endothelial nitric oxide synthase by tetrahydrobiopterin in vascular disease. Arterioscler Thromb Vasc Biol. 2004;24:413–420. doi: 10.1161/01.ATV.0000110785.96039.f6. [DOI] [PubMed] [Google Scholar]
  • [202].Li H, Forstermann U. Uncoupling of endothelial NO synthase in atherosclerosis and vascular disease. Curr Opin Pharmacol. 2013;13:161–167. doi: 10.1016/j.coph.2013.01.006. [DOI] [PubMed] [Google Scholar]
  • [203].Zweier JL, Chen CA, Druhan LJ. S-glutathionylation reshapes our understanding of endothelial nitric oxide synthase uncoupling and nitric oxide/reactive oxygen species-mediated signaling. Antioxid Redox Sig. 2011;14:1769–1775. doi: 10.1089/ars.2011.3904. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [204].Channon KM. Tetrahydrobiopterin: a vascular redox target to improve endothelial function. Curr Vasc Pharmacol. 2012;10:705–708. doi: 10.2174/157016112803520819. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [205].Ding H, Triggle CR. Endothelial dysfunction in diabetes: multiple targets for treatment. Pflugers Archiv. 2010;459:977–994. doi: 10.1007/s00424-010-0807-3. [DOI] [PubMed] [Google Scholar]
  • [206].Chander PN, Gealekman O, Brodsky SV, Elitok S, Tojo A, Crabtree M, Gross SS, Goligorsky MS. Nephropathy in Zucker diabetic fat rat is associated with oxidative and nitrosative stress: prevention by chronic therapy with a peroxynitrite scavenger ebselen. J Am Soc Nephrol. 2004;15:2391–2403. doi: 10.1097/01.ASN.0000135971.88164.2C. [DOI] [PubMed] [Google Scholar]
  • [207].Shinozaki K, Kashiwagi A, Nishio Y, Okamura T, Yoshida Y, Masada M, Toda N, Kikkawa R. Abnormal biopterin metabolism is a major cause of impaired endothelium-dependent relaxation through nitric oxide/O2-imbalance in insulin-resistant rat aorta. Diabetes. 1999;48:2437–2445. doi: 10.2337/diabetes.48.12.2437. [DOI] [PubMed] [Google Scholar]
  • [208].Pannirselvam M, Verma S, Anderson TJ, Triggle CR. Cellular basis of endothelial dysfunction in small mesenteric arteries from spontaneously diabetic (db/db −/−) mice: role of decreased tetrahydrobiopterin bioavailability. Br J Pharmacol. 2002;136:255–263. doi: 10.1038/sj.bjp.0704683. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [209].Cai S, Khoo J, Channon KM. Augmented BH4 by gene transfer restores nitric oxide synthase function in hyperglycemic human endothelial cells. Cardiovasc Res. 2005;65:823–831. doi: 10.1016/j.cardiores.2004.10.040. [DOI] [PubMed] [Google Scholar]
  • [210].Crabtree MJ, Smith CL, Lam G, Goligorsky MS, Gross SS. Ratio of 5,6,7,8-tetrahydrobiopterin to 7,8-dihydrobiopterin in endothelial cells determines glucose-elicited changes in NO vs. superoxide production by eNOS. Am J Physiol Heart Circ Physiol. 2008;294:H1530–1540. doi: 10.1152/ajpheart.00823.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [211].Forstermann U, Li H. Therapeutic effect of enhancing endothelial nitric oxide synthase (eNOS) expression and preventing eNOS uncoupling. Br J Pharmacol. 2011;164:213–223. doi: 10.1111/j.1476-5381.2010.01196.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [212].Kietadisorn R, Juni RP, Moens AL. Tackling endothelial dysfunction by modulating NOS uncoupling: new insights into its pathogenesis and therapeutic possibilities. Am J Physiol Endocrinol Metab. 2012;302:E481–495. doi: 10.1152/ajpendo.00540.2011. [DOI] [PubMed] [Google Scholar]
  • [213].Crabtree MJ, Channon KM. Synthesis and recycling of tetrahydrobiopterin in endothelial function and vascular disease. Nitric Oxide. 2011;25:81–88. doi: 10.1016/j.niox.2011.04.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [214].Sanchez A, Contreras C, Martinez MP, Climent B, Benedito S, Garcia-Sacristan A, Hernandez M, Prieto D. Role of neural NO synthase (nNOS) uncoupling in the dysfunctional nitrergic vasorelaxation of penile arteries from insulin-resistant obese Zucker rats. PLoS One. 2012;7:e36027. doi: 10.1371/journal.pone.0036027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [215].Molnar J, Yu S, Mzhavia N, Pau C, Chereshnev I, Dansky HM. Diabetes induces endothelial dysfunction but does not increase neointimal formation in high-fat diet fed C57BL/6J mice. Circ Res. 2005;96:1178–1184. doi: 10.1161/01.RES.0000168634.74330.ed. [DOI] [PubMed] [Google Scholar]
  • [216].Brodsky SV, Gealekman O, Chen J, Zhang F, Togashi N, Crabtree M, Gross SS, Nasjletti A, Goligorsky MS. Prevention and reversal of premature endothelial cell senescence and vasculopathy in obesity-induced diabetes by ebselen. Circ Res. 2004;94:377–384. doi: 10.1161/01.RES.0000111802.09964.EF. [DOI] [PubMed] [Google Scholar]
  • [217].Cassuto J, Dou H, Czikora I, Szabo A, Patel VS, Kamath V, Belin de Chantemele E, Feher A, Romero MJ, Bagi Z. Peroxynitrite Disrupts Endothelial Caveolae Leading to eNOS Uncoupling and Diminished Flow-Mediated Dilation in Coronary Arterioles of Diabetic Patients. Diabetes. 2013 doi: 10.2337/db13-0577. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [218].Landmesser U, Dikalov S, Price SR, McCann L, Fukai T, Holland SM, Mitch WE, Harrison DG. Oxidation of tetrahydrobiopterin leads to uncoupling of endothelial cell nitric oxide synthase in hypertension. J Clin Invest. 2003;111:1201–1209. doi: 10.1172/JCI14172. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [219].Satoh M, Fujimoto S, Haruna Y, Arakawa S, Horike H, Komai N, Sasaki T, Tsujioka K, Makino H, Kashihara N. NAD(P)H oxidase and uncoupled nitric oxide synthase are major sources of glomerular superoxide in rats with experimental diabetic nephropathy. Am J Physiol Renal Physiol. 2005;288:F1144–1152. doi: 10.1152/ajprenal.00221.2004. [DOI] [PubMed] [Google Scholar]
  • [220].Bitar MS, Wahid S, Mustafa S, Al-Saleh E, Dhaunsi GS, Al-Mulla F. Nitric oxide dynamics and endothelial dysfunction in type II model of genetic diabetes. Eur J Pharmacol. 2005;511:53–64. doi: 10.1016/j.ejphar.2005.01.014. [DOI] [PubMed] [Google Scholar]
  • [221].Xu J, Xie Z, Reece R, Pimental D, Zou MH. Uncoupling of endothelial nitric oxidase synthase by hypochlorous acid: role of NAD(P)H oxidase-derived superoxide and peroxynitrite. Arterioscler Thromb Vasc Biol. 2006;26:2688–2695. doi: 10.1161/01.ATV.0000249394.94588.82. [DOI] [PubMed] [Google Scholar]
  • [222].Dikalova AE, Gongora MC, Harrison DG, Lambeth JD, Dikalov S, Griendling KK. Upregulation of Nox1 in vascular smooth muscle leads to impaired endothelium-dependent relaxation via eNOS uncoupling. Am J Physiol Heart Circ Physiol. 2010;299:H673–679. doi: 10.1152/ajpheart.00242.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [223].Zhang Q, Malik P, Pandey D, Gupta S, Jagnandan D, Belin de Chantemele E, Banfi B, Marrero MB, Rudic RD, Stepp DW, Fulton DJ. Paradoxical activation of endothelial nitric oxide synthase by NADPH oxidase. Arterioscler Thromb Vasc Biol. 2008;28:1627–1633. doi: 10.1161/ATVBAHA.108.168278. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [224].Koppenol WH. 100 years of peroxynitrite chemistry and 11 years of peroxynitrite biochemistry. Redox Rep. 2001;6:339–341. doi: 10.1179/135100001101536517. [DOI] [PubMed] [Google Scholar]
  • [225].Toutouzas K, Riga M, Stefanadi E, Stefanadis C. Asymmetric dimethylarginine (ADMA) and other endogenous nitric oxide synthase (NOS) inhibitors as an important cause of vascular insulin resistance. Horm Metab Res. 2008;40:655–659. doi: 10.1055/s-0028-1083814. [DOI] [PubMed] [Google Scholar]
  • [226].Risbano MG, Gladwin MT. Therapeutics targeting of dysregulated redox equilibrium and endothelial dysfunction. Handb Exp Pharmacol. 2013;218:315–349. doi: 10.1007/978-3-642-38664-0_13. [DOI] [PubMed] [Google Scholar]
  • [227].Chen CA, Wang TY, Varadharaj S, Reyes LA, Hemann C, Talukder MA, Chen YR, Druhan LJ, Zweier JL. S-glutathionylation uncouples eNOS and regulates its cellular and vascular function. Nature. 2010;468:1115–1118. doi: 10.1038/nature09599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [228].Lei H, Luo S, Qin H, Xia Y. Molecular Mechanisms of Endothelial NO Synthase Uncoupling. Curr Pharm Des. 2013 doi: 10.2174/13816128113196660746. [DOI] [PubMed] [Google Scholar]
  • [229].Williams IL, Wheatcroft SB, Shah AM, Kearney MT. Obesity, atherosclerosis and the vascular endothelium: mechanisms of reduced nitric oxide bioavailability in obese humans. Int J Obes Relat Metab Disord. 2002;26:754–764. doi: 10.1038/sj.ijo.0801995. [DOI] [PubMed] [Google Scholar]
  • [230].Steinberg HO, Chaker H, Leaming R, Johnson A, Brechtel G, Baron AD. Obesity/insulin resistance is associated with endothelial dysfunction. Implications for the syndrome of insulin resistance. J Clin Invest. 1996;97:2601–2610. doi: 10.1172/JCI118709. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [231].Laine H, Yki-Jarvinen H, Kirvela O, Tolvanen T, Raitakari M, Solin O, Haaparanta M, Knuuti J, Nuutila P. Insulin resistance of glucose uptake in skeletal muscle cannot be ameliorated by enhancing endothelium-dependent blood flow in obesity. J Clin Invest. 1998;101:1156–1162. doi: 10.1172/JCI1065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [232].Van Guilder GP, Stauffer BL, Greiner JJ, Desouza CA. Impaired endothelium-dependent vasodilation in overweight and obese adult humans is not limited to muscarinic receptor agonists. Am J Physiol Heart Circ Physiol. 2008;294:H1685–1692. doi: 10.1152/ajpheart.01281.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [233].Arcaro G, Zamboni M, Rossi L, Turcato E, Covi G, Armellini F, Bosello O, Lechi A. Body fat distribution predicts the degree of endothelial dysfunction in uncomplicated obesity. Int J Obes Relat Disord. 1999;23:936–942. doi: 10.1038/sj.ijo.0801022. [DOI] [PubMed] [Google Scholar]
  • [234].Tack CJ, Ong MK, Lutterman JA, Smits P. Insulin-induced vasodilatation and endothelial function in obesity/insulin resistance. Effects of troglitazone. Diabetologia. 1998;41:569–576. doi: 10.1007/s001250050948. [DOI] [PubMed] [Google Scholar]
  • [235].Westerbacka J, Vehkavaara S, Bergholm R, Wilkinson I, Cockcroft J, Yki-Jarvinen H. Marked resistance of the ability of insulin to decrease arterial stiffness characterizes human obesity. Diabetes. 1999;48:821–827. doi: 10.2337/diabetes.48.4.821. [DOI] [PubMed] [Google Scholar]
  • [236].Sturm W, Sandhofer A, Engl J, Laimer M, Molnar C, Kaser S, Weiss H, Tilg H, Ebenbichler CF, Patsch JR. Influence of visceral obesity and liver fat on vascular structure and function in obese subjects. Obesity. 2009;17:1783–1788. doi: 10.1038/oby.2009.81. [DOI] [PubMed] [Google Scholar]
  • [237].Parikh NI, Keyes MJ, Larson MG, Pou KM, Hamburg NM, Vita JA, O’Donnell CJ, Vasan RS, Mitchell GF, Hoffmann U, Fox CS, Benjamin EJ. Visceral and subcutaneous adiposity and brachial artery vasodilator function. Obesity. 2009;17:2054–2059. doi: 10.1038/oby.2009.60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [238].Gupta AK, Cornelissen G, Greenway FL, Dhoopati V, Halberg F, Johnson WD. Abnormalities in circadian blood pressure variability and endothelial function: pragmatic markers for adverse cardiometabolic profiles in asymptomatic obese adults. Cardiovasc Diabetol. 2010;9:58. doi: 10.1186/1475-2840-9-58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [239].Grassi G, Seravalle G, Scopelliti F, Dell’Oro R, Fattori L, Quarti-Trevano F, Brambilla G, Schiffrin EL, Mancia G. Structural and functional alterations of subcutaneous small resistance arteries in severe human obesity. Obesity. 2010;18:92–98. doi: 10.1038/oby.2009.195. [DOI] [PubMed] [Google Scholar]
  • [240].Weil BR, Westby CM, Van Guilder GP, Greiner JJ, Stauffer BL, DeSouza CA. Enhanced endothelin-1 system activity with overweight and obesity. Am J Physiol Heart Circ Physiol. 2011;301:H689–695. doi: 10.1152/ajpheart.00206.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [241].Lambert E, Sari CI, Dawood T, Nguyen J, McGrane M, Eikelis N, Chopra R, Wong C, Chatzivlastou K, Head G, Straznicky N, Esler M, Schlaich M, Lambert G. Sympathetic nervous system activity is associated with obesity-induced subclinical organ damage in young adults. Hypertension. 2010;56:351–358. doi: 10.1161/HYPERTENSIONAHA.110.155663. [DOI] [PubMed] [Google Scholar]
  • [242].Han KA, Patel Y, Lteif AA, Chisholm R, Mather KJ. Contributions of dysglycaemia, obesity, and insulin resistance to impaired endothelium-dependent vasodilation in humans. Diabetes Met Res Rev. 2011;27:354–361. doi: 10.1002/dmrr.1183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [243].Andersson J, Sjostrom LG, Karlsson M, Wiklund U, Hultin M, Karpe F, Olsson T. Dysregulation of subcutaneous adipose tissue blood flow in overweight postmenopausal women. Menopause. 2010;17:365–371. doi: 10.1097/gme.0b013e3181c12b26. [DOI] [PubMed] [Google Scholar]
  • [244].Miadi-Messaoud H, Chouchane A, Abderrazek E, Debbabi H, Zaouali-Ajina M, Tabka Z, Ben-Jebria A. Obesity-induced impairment of endothelium-dependent vasodilation in Tunisian women. Int J Obesity. 2010;34:273–279. doi: 10.1038/ijo.2009.231. [DOI] [PubMed] [Google Scholar]
  • [245].Mahmud FH, Hill DJ, Cuerden MS, Clarson CL. Impaired vascular function in obese adolescents with insulin resistance. J Pediatr. 2009;155:678–682. doi: 10.1016/j.jpeds.2009.04.060. [DOI] [PubMed] [Google Scholar]
  • [246].Bhattacharjee R, Alotaibi WH, Kheirandish-Gozal L, Capdevila OS, Gozal D. Endothelial dysfunction in obese non-hypertensive children without evidence of sleep disordered breathing. BMC pediatr. 2010;10:8. doi: 10.1186/1471-2431-10-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [247].Bhattacharjee R, Kim J, Alotaibi WH, Kheirandish-Gozal L, Capdevila OS, Gozal D. Endothelial dysfunction in children without hypertension: potential contributions of obesity and obstructive sleep apnea. Chest. 2012;141:682–691. doi: 10.1378/chest.11-1777. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [248].El Assar M, Ruiz de Adana JC, Angulo J, Pindado Martinez ML, Hernandez Matias A, Rodriguez-Manas L. Preserved endothelial function in human obesity in the absence of insulin resistance. J Transl Med. 2013;11:263. doi: 10.1186/1479-5876-11-263. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [249].Biasucci LM, Graziani F, Rizzello V, Liuzzo G, Guidone C, De Caterina AR, Brugaletta S, Mingrone G, Crea F. Paradoxical preservation of vascular function in severe obesity. Am J Med. 2010;123:727–734. doi: 10.1016/j.amjmed.2010.02.016. [DOI] [PubMed] [Google Scholar]
  • [250].Czernichow S, Greenfield JR, Galan P, Bastard JP, Charnaux N, Samaras K, Safar ME, Blacher J, Hercberg S, Levy BI. Microvascular dysfunction in healthy insulin-sensitive overweight individuals. J Hypertens. 2010;28:325–332. doi: 10.1097/HJH.0b013e328333d1fc. [DOI] [PubMed] [Google Scholar]
  • [251].Morley JE, Flood JF. Evidence that nitric oxide modulates food intake in mice. Life Sci. 1991;49:707–711. doi: 10.1016/0024-3205(91)90102-h. [DOI] [PubMed] [Google Scholar]
  • [252].Morley JE, Flood JF. Competitive antagonism of nitric oxide synthetase causes weight loss in mice. Life Sci. 1992;51:1285–1289. doi: 10.1016/0024-3205(92)90018-k. [DOI] [PubMed] [Google Scholar]
  • [253].Morley JE, Flood JF. Effect of competitive antagonism of NO synthetase on weight and food intake in obese and diabetic mice. Am J Physiol. 1994;266:R164–168. doi: 10.1152/ajpregu.1994.266.1.R164. [DOI] [PubMed] [Google Scholar]
  • [254].Squadrito F, Calapai G, Cucinotta D, Altavilla D, Zingarelli B, Ioculano M, Urna G, Sardella A, Campo GM, Caputi AP. Anorectic activity of NG-nitro-L-arginine, an inhibitor of brain nitric oxide synthase, in obese Zucker rats. Eur J Pharmacol. 1993;230:125–128. doi: 10.1016/0014-2999(93)90422-e. [DOI] [PubMed] [Google Scholar]
  • [255].Calapai G, Corica F, Allegra A, Corsonello A, Sautebin L, De Gregorio T, Di Rosa M, Costantino G, Buemi M, Caputi AP. Effects of intracerebroventricular leptin administration on food intake, body weight gain and diencephalic nitric oxide synthase activity in the mouse. Br J Pharmacol. 1998;125:798–802. doi: 10.1038/sj.bjp.0702121. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [256].Iuras A, Telles MM, Bertoncini CR, Ko GM, de Andrade IS, Silveira VL, Ribeiro EB. Central administration of a nitric oxide precursor abolishes both the hypothalamic serotonin release and the hypophagia induced by interleukin-1beta in obese Zucker rats. Regul Peptides. 2005;124:145–150. doi: 10.1016/j.regpep.2004.07.007. [DOI] [PubMed] [Google Scholar]
  • [257].Squadrito F, Calapai G, Altavilla D, Cucinotta D, Zingarelli B, Arcoraci V, Campo GM, Caputi AP. Central serotoninergic system involvement in the anorexia induced by NG-nitro-L-arginine, an inhibitor of nitric oxide synthase. Eur J Pharmacol. 1994;255:51–55. doi: 10.1016/0014-2999(94)90081-7. [DOI] [PubMed] [Google Scholar]
  • [258].Stricker-Krongrad A, Beck B, Burlet C. Nitric oxide mediates hyperphagia of obese Zucker rats: relation to specific changes in the microstructure of feeding behavior. Life Sci. 1996;58:PL9–15. doi: 10.1016/0024-3205(95)02260-0. [DOI] [PubMed] [Google Scholar]
  • [259].Tsuchiya K, Sakai H, Suzuki N, Iwashima F, Yoshimoto T, Shichiri M, Hirata Y. Chronic blockade of nitric oxide synthesis reduces adiposity and improves insulin resistance in high fat-induced obese mice. Endocrinology. 2007;148:4548–4556. doi: 10.1210/en.2006-1371. [DOI] [PubMed] [Google Scholar]
  • [260].Shankar R, Zhu JS, Ladd B, Henry D, Shen HQ, Baron AD. Central nervous system nitric oxide synthase activity regulates insulin secretion and insulin action. J Clin Invest. 1998;102:1403–1412. doi: 10.1172/JCI3030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [261].Rodin J. Insulin levels, hunger, and food intake: an example of feedback loops in body weight regulation. Health Psychol. 1985;4:1–24. doi: 10.1037//0278-6133.4.1.1. [DOI] [PubMed] [Google Scholar]
  • [262].Rodin J, Wack J, Ferrannini E, DeFronzo RA. Effect of insulin and glucose on feeding behavior. Metabolism. 1985;34:826–831. doi: 10.1016/0026-0495(85)90106-4. [DOI] [PubMed] [Google Scholar]
  • [263].Louis-Sylvestre J. Meal size: role of reflexly induced insulin release. J Auton Nerv Syst. 1984;10:317–324. doi: 10.1016/0165-1838(84)90029-8. [DOI] [PubMed] [Google Scholar]
  • [264].Woods SC, Gibbs J. The regulation of food intake by peptides. Ann N Y Acad Sci. 1989;575:236–243. doi: 10.1111/j.1749-6632.1989.tb53246.x. [DOI] [PubMed] [Google Scholar]
  • [265].Pliquett RU, Fuhrer D, Falk S, Zysset S, von Cramon DY, Stumvoll M. The effects of insulin on the central nervous system--focus on appetite regulation. Horm Metab Res. 2006;38:442–446. doi: 10.1055/s-2006-947840. [DOI] [PubMed] [Google Scholar]
  • [266].Yamada J, Sugimoto Y, Yoshikawa T, Horisaka K. Effects of a nitric oxide synthase inhibitor on 5-HT1A receptor agonist 8-OH-DPAT-induced hyperphagia in rats. Eur J Pharmacol. 1996;316:23–26. doi: 10.1016/s0014-2999(96)00657-7. [DOI] [PubMed] [Google Scholar]
  • [267].Sugimoto Y, Yamada J, Yoshikawa T. A neuronal nitric oxide synthase inhibitor 7-nitroindazole reduces the 5-HT1A receptor against 8-OH-DPAT-elicited hyperphagia in rats. Eur J Pharmacol. 1999;376:1–5. doi: 10.1016/s0014-2999(99)00378-7. [DOI] [PubMed] [Google Scholar]
  • [268].Sugimoto Y, Yoshikawa T, Yamada J. Involvement of nitric oxide in the 5-HT1A autoreceptor-mediated hyperphagia in rats. Adv Exp Med Biol. 1999;467:109–111. doi: 10.1007/978-1-4615-4709-9_16. [DOI] [PubMed] [Google Scholar]
  • [269].Czech DA, Kazel MR, Harris J. A nitric oxide synthase inhibitor, N(G)-nitro-L-arginine methyl ester, attenuates lipoprivic feeding in mice. Physiol Behav. 2003;80:75–79. doi: 10.1016/s0031-9384(03)00220-8. [DOI] [PubMed] [Google Scholar]
  • [270].Li M, Vizzard MA, Jaworski DM, Galbraith RA. The weight loss elicited by cobalt protoporphyrin is related to decreased activity of nitric oxide synthase in the hypothalamus. J Appl Physiol (1985) 2006;100:1983–1991. doi: 10.1152/japplphysiol.01169.2005. [DOI] [PubMed] [Google Scholar]
  • [271].Wascher TC, Graier WF, Dittrich P, Hussain MA, Bahadori B, Wallner S, Toplak H. Effects of low-dose L-arginine on insulin-mediated vasodilatation and insulin sensitivity. Eur J Clin Invest. 1997;27:690–695. doi: 10.1046/j.1365-2362.1997.1730718.x. [DOI] [PubMed] [Google Scholar]
  • [272].Lucotti P, Setola E, Monti LD, Galluccio E, Costa S, Sandoli EP, Fermo I, Rabaiotti G, Gatti R, Piatti P. Beneficial effects of a long-term oral L-arginine treatment added to a hypocaloric diet and exercise training program in obese, insulin-resistant type 2 diabetic patients. Am J Physiol Endocrinol Metab. 2006;291:E906–912. doi: 10.1152/ajpendo.00002.2006. [DOI] [PubMed] [Google Scholar]
  • [273].Alizadeh M, Safaeiyan A, Ostadrahimi A, Estakhri R, Daneghian S, Ghaffari A, Gargari BP. Effect of L-arginine and selenium added to a hypocaloric diet enriched with legumes on cardiovascular disease risk factors in women with central obesity: a randomized, double-blind, placebo-controlled trial. Ann Nutr Metab. 2012;60:157–168. doi: 10.1159/000335470. [DOI] [PubMed] [Google Scholar]
  • [274].Bogdanski P, Suliburska J, Grabanska K, Musialik K, Cieslewicz A, Skoluda A, Jablecka A. Effect of 3-month L-arginine supplementation on insulin resistance and tumor necrosis factor activity in patients with visceral obesity. Eur Rev Med Sci. 2012;16:816–823. [PubMed] [Google Scholar]
  • [275].Bogdanski P, Szulinska M, Suliburska J, Pupek-Musialik D, Jablecka A, Witmanowski H. Supplementation with L-arginine favorably influences plasminogen activator inhibitor type 1 concentration in obese patients. A randomized, double blind trial. J Endocrinol Invest. 2013;36:221–226. doi: 10.3275/8467. [DOI] [PubMed] [Google Scholar]
  • [276].Suliburska J, Bogdanski P, Szulinska M, Pupek-Musialik D, Jablecka A. Changes in mineral status are associated with improvements in insulin sensitivity in obese patients following L-arginine supplementation. Eur J Nutr. 2014;53:387–393. doi: 10.1007/s00394-013-0533-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [277].Monti LD, Casiraghi MC, Setola E, Galluccio E, Pagani MA, Quaglia L, Bosi E, Piatti P. L-arginine enriched biscuits improve endothelial function and glucose metabolism: a pilot study in healthy subjects and a cross-over study in subjects with impaired glucose tolerance and metabolic syndrome. Metabolism. 2013;62:255–264. doi: 10.1016/j.metabol.2012.08.004. [DOI] [PubMed] [Google Scholar]
  • [278].Clemmensen C, Madsen AN, Smajilovic S, Holst B, Brauner-Osborne H. L-Arginine improves multiple physiological parameters in mice exposed to diet-induced metabolic disturbances. Amino Acids. 2012;43:1265–1275. doi: 10.1007/s00726-011-1199-1. [DOI] [PubMed] [Google Scholar]
  • [279].Fu WJ, Haynes TE, Kohli R, Hu J, Shi W, Spencer TE, Carroll RJ, Meininger CJ, Wu G. Dietary L-arginine supplementation reduces fat mass in Zucker diabetic fatty rats. J Nutr. 2005;135:714–721. doi: 10.1093/jn/135.4.714. [DOI] [PubMed] [Google Scholar]
  • [280].Jobgen W, Meininger CJ, Jobgen SC, Li P, Lee MJ, Smith SB, Spencer TE, Fried SK, Wu G. Dietary L-arginine supplementation reduces white fat gain and enhances skeletal muscle and brown fat masses in diet-induced obese rats. J Nutr. 2009;139:230–237. doi: 10.3945/jn.108.096362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [281].Tan B, Yin Y, Liu Z, Li X, Xu H, Kong X, Huang R, Tang W, Shinzato I, Smith SB, Wu G. Dietary L-arginine supplementation increases muscle gain and reduces body fat mass in growing-finishing pigs. Amino Acids. 2009;37:169–175. doi: 10.1007/s00726-008-0148-0. [DOI] [PubMed] [Google Scholar]
  • [282].McKnight JR, Satterfield MC, Jobgen WS, Smith SB, Spencer TE, Meininger CJ, McNeal CJ, Wu G. Beneficial effects of L-arginine on reducing obesity: potential mechanisms and important implications for human health. Amino Acids. 2010;39:349–357. doi: 10.1007/s00726-010-0598-z. [DOI] [PubMed] [Google Scholar]
  • [283].Ayala JE, Bracy DP, Julien BM, Rottman JN, Fueger PT, Wasserman DH. Chronic treatment with sildenafil improves energy balance and insulin action in high fat-fed conscious mice. Diabetes. 2007;56:1025–1033. doi: 10.2337/db06-0883. [DOI] [PubMed] [Google Scholar]
  • [284].Mitschke MM, Hoffmann LS, Gnad T, Scholz D, Kruithoff K, Mayer P, Haas B, Sassmann A, Pfeifer A, Kilic A. Increased cGMP promotes healthy expansion and browning of white adipose tissue. FASEB J. 2013;27:1621–1630. doi: 10.1096/fj.12-221580. [DOI] [PubMed] [Google Scholar]
  • [285].De Toni L, Strapazzon G, Gianesello L, Caretta N, Pilon C, Bruttocao A, Foresta C. Effects of type 5-phosphodiesterase inhibition on energy metabolism and mitochondrial biogenesis in human adipose tissue ex vivo. J Endocrinol Invest. 2011;34:738–741. doi: 10.1007/BF03346724. [DOI] [PubMed] [Google Scholar]
  • [286].Wang H, Wang AX, Aylor K, Barrett EJ. Nitric oxide directly promotes vascular endothelial insulin transport. Diabetes. 2013;62:4030–4042. doi: 10.2337/db13-0627. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [287].Baron AD, Zhu JS, Marshall S, Irsula O, Brechtel G, Keech C. Insulin resistance after hypertension induced by the nitric oxide synthesis inhibitor L-NMMA in rats. Am J Physiol. 1995;269:E709–715. doi: 10.1152/ajpendo.1995.269.4.E709. [DOI] [PubMed] [Google Scholar]
  • [288].Meshkani R, Adeli K. Hepatic insulin resistance, metabolic syndrome and cardiovascular disease. Clin Biochem. 2009;42:1331–1346. doi: 10.1016/j.clinbiochem.2009.05.018. [DOI] [PubMed] [Google Scholar]
  • [289].Meininger CJ, Marinos RS, Hatakeyama K, Martinez-Zaguilan R, Rojas JD, Kelly KA, Wu G. Impaired nitric oxide production in coronary endothelial cells of the spontaneously diabetic BB rat is due to tetrahydrobiopterin deficiency. Biochem J. 2000;349:353–356. doi: 10.1042/0264-6021:3490353. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [290].Xu J, Wu Y, Song P, Zhang M, Wang S, Zou MH. Proteasome-dependent degradation of guanosine 5′-triphosphate cyclohydrolase I causes tetrahydrobiopterin deficiency in diabetes mellitus. Circulation. 2007;116:944–953. doi: 10.1161/CIRCULATIONAHA.106.684795. [DOI] [PubMed] [Google Scholar]
  • [291].Meininger CJ, Cai S, Parker JL, Channon KM, Kelly KA, Becker EJ, Wood MK, Wade LA, Wu G. GTP cyclohydrolase I gene transfer reverses tetrahydrobiopterin deficiency and increases nitric oxide synthesis in endothelial cells and isolated vessels from diabetic rats. FASEB J. 2004;18:1900–1902. doi: 10.1096/fj.04-1702fje. [DOI] [PubMed] [Google Scholar]
  • [292].Abudukadier A, Fujita Y, Obara A, Ohashi A, Fukushima T, Sato Y, Ogura M, Nakamura Y, Fujimoto S, Hosokawa M, Hasegawa H, Inagaki N. Tetrahydrobiopterin has a glucose-lowering effect by suppressing hepatic gluconeogenesis in an endothelial nitric oxide synthase-dependent manner in diabetic mice. Diabetes. 2013;62:3033–3043. doi: 10.2337/db12-1242. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [293].Sadri P, Lautt WW. Blockade of hepatic nitric oxide synthase causes insulin resistance. Am J Physiol. 1999;277:G101–108. doi: 10.1152/ajpgi.1999.277.1.G101. [DOI] [PubMed] [Google Scholar]
  • [294].Guarino MP, Afonso RA, Raimundo N, Raposo JF, Macedo MP. Hepatic glutathione and nitric oxide are critical for hepatic insulin-sensitizing substance action. Am J Physiol Gastrointest Liv Physiol. 2003;284:G588–594. doi: 10.1152/ajpgi.00423.2002. [DOI] [PubMed] [Google Scholar]
  • [295].Guarino MP, Macedo MP. Co-administration of glutathione and nitric oxide enhances insulin sensitivity in Wistar rats. Br J Pharmacol. 2006;147:959–965. doi: 10.1038/sj.bjp.0706691. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [296].Sosa I, Cvijanovic O, Celic T, Cuculic D, Crncevic-Orlic Z, Vukelic L, Zoricic Cvek S, Dudaric L, Bosnar A, Bobinac D. Hepatoregenerative role of bone morphogenetic protein-9. Med Sci Monit. 2011;17:HY33–35. doi: 10.12659/MSM.882108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [297].Lautt WW. A new paradigm for diabetes and obesity: the hepatic insulin sensitizing substance (HISS) hypothesis. J Pharmacol Sci. 2004;95:9–17. doi: 10.1254/jphs.95.9. [DOI] [PubMed] [Google Scholar]
  • [298].Lautt WW. Insulin sensitivity in skeletal muscle regulated by a hepatic hormone, HISS. Can J Appl Physiol. 2005;30:304–312. doi: 10.1139/h05-123. [DOI] [PubMed] [Google Scholar]
  • [299].Ming Z, Lautt WW. HISS, not insulin, causes vasodilation in response to administered insulin. J Appl Physiol (1985) 2011;110:60–68. doi: 10.1152/japplphysiol.00714.2010. [DOI] [PubMed] [Google Scholar]
  • [300].Steinberg HO, Brechtel G, Johnson A, Fineberg N, Baron AD. Insulin-mediated skeletal muscle vasodilation is nitric oxide dependent. A novel action of insulin to increase nitric oxide release. J Clin Invest. 1994;94:1172–1179. doi: 10.1172/JCI117433. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [301].Scherrer U, Randin D, Vollenweider P, Vollenweider L, Nicod P. Nitric oxide release accounts for insulin’s vascular effects in humans. J Clin Invest. 1994;94:2511–2515. doi: 10.1172/JCI117621. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [302].Baron AD. Hemodynamic actions of insulin. Am J Physiol. 1994;267:E187–202. doi: 10.1152/ajpendo.1994.267.2.E187. [DOI] [PubMed] [Google Scholar]
  • [303].Guarino MP, Correia NC, Lautt WW, Macedo MP. Insulin sensitivity is mediated by the activation of the ACh/NO/cGMP pathway in rat liver. Am J Physiol Gastrointest Liv Physiol. 2004;287:G527–532. doi: 10.1152/ajpgi.00085.2004. [DOI] [PubMed] [Google Scholar]
  • [304].Guarino MP, Correia NC, Raposo J, Macedo MP. Nitric oxide synthase inhibition decreases output of hepatic insulin sensitizing substance (HISS), which is reversed by SIN-1 but not by nitroprusside. Proc West Pharmacol Soc. 2001;44:25–26. [PubMed] [Google Scholar]
  • [305].Lautt WW, Macedo MP, Sadri P, Takayama S, Duarte Ramos F, Legare DJ. Hepatic parasympathetic (HISS) control of insulin sensitivity determined by feeding and fasting. Am J Physiol Gastrointest Liv Physiol. 2001;281:G29–36. doi: 10.1152/ajpgi.2001.281.1.G29. [DOI] [PubMed] [Google Scholar]
  • [306].Barreto SG, Carati CJ, Toouli J, Saccone GT. The islet-acinar axis of the pancreas: more than just insulin. Am J Physiol Gastrointest Liv Physiol. 2010;299:G10–22. doi: 10.1152/ajpgi.00077.2010. [DOI] [PubMed] [Google Scholar]
  • [307].Chey WY, Chang T. Neural hormonal regulation of exocrine pancreatic secretion. Pancreatology. 2001;1:320–335. doi: 10.1159/000055831. [DOI] [PubMed] [Google Scholar]
  • [308].Vincent SR. Nitric oxide and arginine-evoked insulin secretion. Science. 1992;258:1376–1378. doi: 10.1126/science.1455235. [DOI] [PubMed] [Google Scholar]
  • [309].Broniowska KA, Oleson BJ, Corbett JA. beta-Cell Responses to Nitric Oxide. Vitam Horm. 2014;95:299–322. doi: 10.1016/B978-0-12-800174-5.00012-0. [DOI] [PubMed] [Google Scholar]
  • [310].Takamura T, Kato I, Kimura N, Nakazawa T, Yonekura H, Takasawa S, Okamoto H. Transgenic mice overexpressing type 2 nitric-oxide synthase in pancreatic beta cells develop insulin-dependent diabetes without insulitis. J Biol Chem. 1998;273:2493–2496. doi: 10.1074/jbc.273.5.2493. [DOI] [PubMed] [Google Scholar]
  • [311].Laychock SG, Modica ME, Cavanaugh CT. L-arginine stimulates cyclic guanosine 3′,5′-monophosphate formation in rat islets of Langerhans and RINm5F insulinoma cells: evidence for L-arginine:nitric oxide synthase. Endocrinology. 1991;129:3043–3052. doi: 10.1210/endo-129-6-3043. [DOI] [PubMed] [Google Scholar]
  • [312].Schmidt HH, Warner TD, Ishii K, Sheng H, Murad F. Insulin secretion from pancreatic B cells caused by L-arginine-derived nitrogen oxides. Science. 1992;255:721–723. doi: 10.1126/science.1371193. [DOI] [PubMed] [Google Scholar]
  • [313].Spinas GA, Laffranchi R, Francoys I, David I, Richter C, Reinecke M. The early phase of glucose-stimulated insulin secretion requires nitric oxide. Diabetologia. 1998;41:292–299. doi: 10.1007/s001250050906. [DOI] [PubMed] [Google Scholar]
  • [314].Spinas GA. The Dual Role of Nitric Oxide in Islet beta-Cells. News Physiol Sci. 1999;14:49–54. doi: 10.1152/physiologyonline.1999.14.2.49. [DOI] [PubMed] [Google Scholar]
  • [315].Laffranchi R, Gogvadze V, Richter C, Spinas GA. Nitric oxide (nitrogen monoxide, NO) stimulates insulin secretion by inducing calcium release from mitochondria. Biochem Biophys Res Comm. 1995;217:584–591. doi: 10.1006/bbrc.1995.2815. [DOI] [PubMed] [Google Scholar]
  • [316].Jones PM, Persaud SJ, Bjaaland T, Pearson JD, Howell SL. Nitric oxide is not involved in the initiation of insulin secretion from rat islets of Langerhans. Diabetologia. 1992;35:1020–1027. doi: 10.1007/BF02221676. [DOI] [PubMed] [Google Scholar]
  • [317].Weigert N, Dollinger M, Schmid R, Schusdziarra V. Contribution of neural intrapancreatic non-cholinergic non-adrenergic mechanisms to glucose-induced insulin release in the isolated rat pancreas. Diabetologia. 1992;35:1133–1139. doi: 10.1007/BF00401366. [DOI] [PubMed] [Google Scholar]
  • [318].Ohtoshi K, Yamasaki Y, Gorogawa S, Hayaishi-Okano R, Node K, Matsuhisa M, Kajimoto Y, Hori M. Association of (−)786T-C mutation of endothelial nitric oxide synthase gene with insulin resistance. Diabetologia. 2002;45:1594–1601. doi: 10.1007/s00125-002-0922-6. [DOI] [PubMed] [Google Scholar]
  • [319].Yoshimura T, Hisatomi A, Kajihara S, Yasutake T, Ogawa Y, Mizuta T, Ozaki I, Utsunomiyai T, Yamamoto K. The relationship between insulin resistance and polymorphisms of the endothelial nitric oxide synthase gene in patients with coronary artery disease. J Atheroscler Thromb. 2003;10:43–47. doi: 10.5551/jat.10.43. [DOI] [PubMed] [Google Scholar]
  • [320].Vecoli C, Andreassi MG, Liga R, Colombo MG, Coceani M, Carpeggiani C, L’Abbate A, Neglia D. T(−786)-->C polymorphism of the endothelial nitric oxide synthase gene is associated with insulin resistance in patients with ischemic or non ischemic cardiomyopathy. BMC Med Genet. 2012;13:92. doi: 10.1186/1471-2350-13-92. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [321].Monti LD, Barlassina C, Citterio L, Galluccio E, Berzuini C, Setola E, Valsecchi G, Lucotti P, Pozza G, Bernardinelli L, Casari G, Piatti P. Endothelial nitric oxide synthase polymorphisms are associated with type 2 diabetes and the insulin resistance syndrome. Diabetes. 2003;52:1270–1275. doi: 10.2337/diabetes.52.5.1270. [DOI] [PubMed] [Google Scholar]
  • [322].Gonzalez-Sanchez JL, Martinez-Larrad MT, Saez ME, Zabena C, Martinez-Calatrava MJ, Serrano-Rios M. Endothelial nitric oxide synthase haplotypes are associated with features of metabolic syndrome. Clin Chem. 2007;53:91–97. doi: 10.1373/clinchem.2006.075176. [DOI] [PubMed] [Google Scholar]
  • [323].Rittig K, Holder K, Stock J, Tschritter O, Peter A, Stefan N, Fritsche A, Machicao F, Haring HU, Balletshofer B. Endothelial NO-synthase intron 4 polymorphism is associated with disturbed in vivo nitric oxide production in individuals prone to type 2 diabetes. Horm Metab Res. 2008;40:13–17. doi: 10.1055/s-2007-1004527. [DOI] [PubMed] [Google Scholar]
  • [324].Morris BJ, Glenn CL, Wilcken DE, Wang XL. Influence of an inducible nitric oxide synthase promoter variant on clinical variables in patients with coronary artery disease. Clin Sci. 2001;100:551–556. [PubMed] [Google Scholar]
  • [325].Bagarolli RA, Saad MJ, Saad ST. Toll-like receptor 4 and inducible nitric oxide synthase gene polymorphisms are associated with Type 2 diabetes. Journal of diabetes and its complications. 2010;24:192–198. doi: 10.1016/j.jdiacomp.2009.03.003. [DOI] [PubMed] [Google Scholar]
  • [326].Duplain H, Burcelin R, Sartori C, Cook S, Egli M, Lepori M, Vollenweider P, Pedrazzini T, Nicod P, Thorens B, Scherrer U. Insulin resistance, hyperlipidemia, and hypertension in mice lacking endothelial nitric oxide synthase. Circulation. 2001;104:342–345. doi: 10.1161/01.cir.104.3.342. [DOI] [PubMed] [Google Scholar]
  • [327].Cook S, Hugli O, Egli M, Menard B, Thalmann S, Sartori C, Perrin C, Nicod P, Thorens B, Vollenweider P, Scherrer U, Burcelin R. Partial gene deletion of endothelial nitric oxide synthase predisposes to exaggerated high-fat diet-induced insulin resistance and arterial hypertension. Diabetes. 2004;53:2067–2072. doi: 10.2337/diabetes.53.8.2067. [DOI] [PubMed] [Google Scholar]
  • [328].Cook S, Hugli O, Egli M, Vollenweider P, Burcelin R, Nicod P, Thorens B, Scherrer U. Clustering of cardiovascular risk factors mimicking the human metabolic syndrome X in eNOS null mice. Swiss Med Wkly. 2003;133:360–363. doi: 10.4414/smw.2003.10239. [DOI] [PubMed] [Google Scholar]
  • [329].Shankar RR, Wu Y, Shen HQ, Zhu JS, Baron AD. Mice with gene disruption of both endothelial and neuronal nitric oxide synthase exhibit insulin resistance. Diabetes. 2000;49:684–687. doi: 10.2337/diabetes.49.5.684. [DOI] [PubMed] [Google Scholar]
  • [330].Nakata S, Tsutsui M, Shimokawa H, Suda O, Morishita T, Shibata K, Yatera Y, Sabanai K, Tanimoto A, Nagasaki M, Tasaki H, Sasaguri Y, Nakashima Y, Otsuji Y, Yanagihara N. Spontaneous myocardial infarction in mice lacking all nitric oxide synthase isoforms. Circulation. 2008;117:2211–2223. doi: 10.1161/CIRCULATIONAHA.107.742692. [DOI] [PubMed] [Google Scholar]
  • [331].Sydow K, Mondon CE, Schrader J, Konishi H, Cooke JP. Dimethylarginine dimethylaminohydrolase overexpression enhances insulin sensitivity. Arterioscler Thromb Vasc Biol. 2008;28:692–697. doi: 10.1161/ATVBAHA.108.162073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [332].Le Gouill E, Jimenez M, Binnert C, Jayet PY, Thalmann S, Nicod P, Scherrer U, Vollenweider P. Endothelial nitric oxide synthase (eNOS) knockout mice have defective mitochondrial beta-oxidation. Diabetes. 2007;56:2690–2696. doi: 10.2337/db06-1228. [DOI] [PubMed] [Google Scholar]
  • [333].Lee-Young RS, Ayala JE, Hunley CF, James FD, Bracy DP, Kang L, Wasserman DH. Endothelial nitric oxide synthase is central to skeletal muscle metabolic regulation and enzymatic signaling during exercise in vivo. Am J Physiol Regul Integr Comp Physiol. 2010;298:R1399–1408. doi: 10.1152/ajpregu.00004.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [334].Carlstrom M, Larsen FJ, Nystrom T, Hezel M, Borniquel S, Weitzberg E, Lundberg JO. Dietary inorganic nitrate reverses features of metabolic syndrome in endothelial nitric oxide synthase-deficient mice. Proc Natl Acad Sci U S A. 2010;107:17716–17720. doi: 10.1073/pnas.1008872107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [335].Sansbury BE, Cummins TD, Tang Y, Hellmann J, Holden CR, Harbeson MA, Chen Y, Patel RP, Spite M, Bhatnagar A, Hill BG. Overexpression of endothelial nitric oxide synthase prevents diet-induced obesity and regulates adipocyte phenotype. Circ Res. 2012;111:1176–1189. doi: 10.1161/CIRCRESAHA.112.266395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [336].Atochin DN, Huang PL. Endothelial nitric oxide synthase transgenic models of endothelial dysfunction. Pflugers Archiv. 2010;460:965–974. doi: 10.1007/s00424-010-0867-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [337].Kashiwagi S, Atochin DN, Li Q, Schleicher M, Pong T, Sessa WC, Huang PL. eNOS phosphorylation on serine 1176 affects insulin sensitivity and adiposity. Biochem Biophys Res Comm. 2013;431:284–290. doi: 10.1016/j.bbrc.2012.12.110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [338].Khedara A, Kawai Y, Kayashita J, Kato N. Feeding rats the nitric oxide synthase inhibitor, L-N(omega)nitroarginine, elevates serum triglyceride and cholesterol and lowers hepatic fatty acid oxidation. J Nutr. 1996;126:2563–2567. doi: 10.1093/jn/126.10.2563. [DOI] [PubMed] [Google Scholar]
  • [339].Khedara A, Goto T, Morishima M, Kayashita J, Kato N. Elevated body fat in rats by the dietary nitric oxide synthase inhibitor, L-N omega nitroarginine. Biosci Biotechnol Biochem. 1999;63:698–702. doi: 10.1271/bbb.63.698. [DOI] [PubMed] [Google Scholar]
  • [340].Recchia FA, Osorio JC, Chandler MP, Xu X, Panchal AR, Lopaschuk GD, Hintze TH, Stanley WC. Reduced synthesis of NO causes marked alterations in myocardial substrate metabolism in conscious dogs. Am J Physiol Endocrinol Metab. 2002;282:E197–206. doi: 10.1152/ajpendo.2002.282.1.E197. [DOI] [PubMed] [Google Scholar]
  • [341].Garcia-Villafranca J, Guillen A, Castro J. Involvement of nitric oxide/cyclic GMP signaling pathway in the regulation of fatty acid metabolism in rat hepatocytes. Biochem Pharmacol. 2003;65:807–812. doi: 10.1016/s0006-2952(02)01623-4. [DOI] [PubMed] [Google Scholar]
  • [342].Goto T, Ohnomi S, Khedara A, Kato N, Ogawa H, Yanagita T. Feeding the nitric oxide synthase inhibitor L-N(omega)nitroarginine elevates serum very low density lipoprotein and hepatic triglyceride synthesis in rats. J Nutr Biochem. 1999;10:274–278. doi: 10.1016/s0955-2863(99)00008-x. [DOI] [PubMed] [Google Scholar]
  • [343].Schild L, Dombrowski F, Lendeckel U, Schulz C, Gardemann A, Keilhoff G. Impairment of endothelial nitric oxide synthase causes abnormal fat and glycogen deposition in liver. Biochim Biophys Acta. 2008;1782:180–187. doi: 10.1016/j.bbadis.2007.12.007. [DOI] [PubMed] [Google Scholar]
  • [344].Lefebvre P, Chinetti G, Fruchart JC, Staels B. Sorting out the roles of PPAR alpha in energy metabolism and vascular homeostasis. J Clin Invest. 2006;116:571–580. doi: 10.1172/JCI27989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [345].Doulias PT, Tenopoulou M, Greene JL, Raju K, Ischiropoulos H. Nitric oxide regulates mitochondrial fatty acid metabolism through reversible protein S-nitrosylation. Sci Signal. 2013;6(256):rs1. doi: 10.1126/scisignal.2003252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [346].Perreault M, Marette A. Targeted disruption of inducible nitric oxide synthase protects against obesity-linked insulin resistance in muscle. Nat Med. 2001;7:1138–1143. doi: 10.1038/nm1001-1138. [DOI] [PubMed] [Google Scholar]
  • [347].Carvalho-Filho MA, Ueno M, Hirabara SM, Seabra AB, Carvalheira JB, de Oliveira MG, Velloso LA, Curi R, Saad MJ. S-nitrosation of the insulin receptor, insulin receptor substrate 1, and protein kinase B/Akt: a novel mechanism of insulin resistance. Diabetes. 2005;54:959–967. doi: 10.2337/diabetes.54.4.959. [DOI] [PubMed] [Google Scholar]
  • [348].Carvalho-Filho MA, Ueno M, Carvalheira JB, Velloso LA, Saad MJ. Targeted disruption of iNOS prevents LPS-induced S-nitrosation of IRbeta/IRS-1 and Akt and insulin resistance in muscle of mice. Am J Physiol Endocrinol Metab. 2006;291:E476–482. doi: 10.1152/ajpendo.00422.2005. [DOI] [PubMed] [Google Scholar]
  • [349].Pauli JR, Ropelle ER, Cintra DE, Carvalho-Filho MA, Moraes JC, De Souza CT, Velloso LA, Carvalheira JB, Saad MJ. Acute physical exercise reverses S-nitrosation of the insulin receptor, insulin receptor substrate 1 and protein kinase B/Akt in diet-induced obese Wistar rats. J Physiol. 2008;586:659–671. doi: 10.1113/jphysiol.2007.142414. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [350].Williamson RT. On the Treatment of Glycosuria and Diabetes Mellitus with Sodium Salicylate. Br Med J. 1901;1:760–762. doi: 10.1136/bmj.1.2100.760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [351].Smith MJ, Meade BW. The effect of salicylate on glycosuria, blood glucose and liver glycogen of the alloxan-diabetic rat. Biochem J. 1952;51:18–20. doi: 10.1042/bj0510018. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [352].Hundal RS, Petersen KF, Mayerson AB, Randhawa PS, Inzucchi S, Shoelson SE, Shulman GI. Mechanism by which high-dose aspirin improves glucose metabolism in type 2 diabetes. J Clin Invest. 2002;109:1321–1326. doi: 10.1172/JCI14955. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [353].Kim JK, Kim YJ, Fillmore JJ, Chen Y, Moore I, Lee J, Yuan M, Li ZW, Karin M, Perret P, Shoelson SE, Shulman GI. Prevention of fat-induced insulin resistance by salicylate. J Clin Invest. 2001;108:437–446. doi: 10.1172/JCI11559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [354].Carvalho-Filho MA, Ropelle ER, Pauli RJ, Cintra DE, Tsukumo DM, Silveira LR, Curi R, Carvalheira JB, Velloso LA, Saad MJ. Aspirin attenuates insulin resistance in muscle of diet-induced obese rats by inhibiting inducible nitric oxide synthase production and S-nitrosylation of IRbeta/IRS-1 and Akt. Diabetologia. 2009;52:2425–2434. doi: 10.1007/s00125-009-1498-1. [DOI] [PubMed] [Google Scholar]
  • [355].Shinozaki S, Choi CS, Shimizu N, Yamada M, Kim M, Zhang T, Shiota G, Dong HH, Kim YB, Kaneki M. Liver-specific inducible nitric-oxide synthase expression is sufficient to cause hepatic insulin resistance and mild hyperglycemia in mice. J Biol Chem. 2011;286:34959–34975. doi: 10.1074/jbc.M110.187666. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [356].Handa P, Tateya S, Rizzo NO, Cheng AM, Morgan-Stevenson V, Han CY, Clowes AW, Daum G, O’Brien KD, Schwartz MW, Chait A, Kim F. Reduced vascular nitric oxide-cGMP signaling contributes to adipose tissue inflammation during high-fat feeding. Arterioscler Thromb Vasc Biol. 2011;31:2827–2835. doi: 10.1161/ATVBAHA.111.236554. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [357].Cha HN, Kim YW, Kim JY, Kim YD, Song IH, Min KN, Park SY. Lack of inducible nitric oxide synthase does not prevent aging-associated insulin resistance. Exp Gerontol. 2010;45:711–718. doi: 10.1016/j.exger.2010.05.004. [DOI] [PubMed] [Google Scholar]
  • [358].Taniguchi CM, Emanuelli B, Kahn CR. Critical nodes in signalling pathways: insights into insulin action. Nat Rev Mol Cell Biol. 2006;7:85–96. doi: 10.1038/nrm1837. [DOI] [PubMed] [Google Scholar]
  • [359].Turini P, Thalmann S, Jayet PY, Cook S, Mathieu C, Burcelin R, Nicod P, Vollenweider P, Sartori C, Scherrer U. Insulin resistance in mice lacking neuronal nitric oxide synthase is related to an alpha-adrenergic mechanism. Swiss Med Wkly. 2007;137:700–704. doi: 10.4414/smw.2007.11950. [DOI] [PubMed] [Google Scholar]

RESOURCES