Abstract
Necrotizing Enterocolitis (NEC) is the most common life-threatening gastrointestinal disease encountered in the premature infant. Although the inciting events leading to NEC remain elusive, various risk factors including prematurity, hypoxemia, formula feeding and intestinal ischemia have been implicated in the pathogenesis of NEC. Data from our lab and others suggest that NEC evolves from disruption of the intestinal epithelial barrier, as a result of a combination of local and systemic insults. We postulate that nitric oxide (NO), an important second messenger and inflammatory mediator, plays a key role in intestinal barrier failure seen in NEC. Nitric oxide and its reactive nitrogen derivative, peroxynitrite, may affect gut barrier permeability by inducing enterocyte apoptosis (programmed cell death) and necrosis, or by disrupting tight junctions or gap junctions that normally play a key role in maintaining epithelial monolayer integrity. Intrinsic mechanisms that serve to restore monolayer integrity following epithelial injury include enterocyte proliferation, epithelial restitution via enterocyte migration, and re-establishment of cell contacts. This review focuses on the biology of NO and the mechanisms by which it promotes epithelial injury while concurrently disrupting the intrinsic repair mechanisms.
INDEX WORDS: Necrotizing Enterocolitis, Nitric Oxide, Intestinal inflammation, Intestinal restitution
Necrotizing Enterocolitis (NEC) is the most common life-threatening gastrointestinal disease encountered in the premature infant. The incidence of NEC approaches 1 per 1000 live-births, with approximately 1 out of 7 affected neonates succumbing to the disease1. The mortality rate for the most severe form of the disease, which is characterized by involvement of the entire intestine, approaches 100%2. Although the inciting events leading to NEC remain elusive, various risk factors including prematurity, hypoxemia, formula feeding and intestinal ischemia have been implicated in the pathogenesis of NEC. Advances in neonatology over the last few decades have increased survival of very low birth weight infants. As a result, the population of patients at risk for developing NEC is increasing. Given the decreased mortality from respiratory distress syndrome3, current trends predict that NEC may soon become the leading cause of death in premature infants4. Because of the increasing incidence of NEC and the associated mortality and long-term complications seen in surviving patients, the primary focus of research in NEC has shifted to the prevention and treatment of the disease.
Our group and others have shown that the pathogenesis of NEC involves disruption of the intestinal epithelial barrier, due to a combination of local and systemic insults5. The intact intestinal barrier comprises several layers of defense against bacterial invasion such as the mucus layer, the epithelium, and the underlying mucosal immune system. The intestinal epithelial barrier promotes immune surveillance against intraluminal microorganisms and macromolecules, by facilitating the continuous sampling of luminal antigens. However, pathologic or dysregulated transfer of antigens across the intestinal barrier can lead to various diseases, including inflammatory bowel disease in adults, and NEC in premature infants6, 7. Intestinal epithelial cells function as both antigen processing and immune effector cells, producing pro-inflammatory cytokines and other mediators when stimulated8. Clinical studies have shown that inflammatory mediators, such as TNFα, IL-1, platelet activating factor, and nitric oxide (NO), produced by enterocytes and macrophages, contribute to the pathogenesis of NEC9–11.
We postulate that NO, an important second messenger and inflammatory mediator, plays a key role in intestinal barrier failure seen in NEC. This review focuses on the biology of NO and the mechanisms by which it mediates intestinal epithelial injury.
Biology of NO
NO is a short-lived reactive molecule that plays an important role in many physiological and pathophysiological processes. It was first described as endothelial derived relaxing factor (EDRF); a potent, short-lived vasodilator. Subsequent studies revealed that in addition to its role as a regulator of blood vessel tone, NO also modulates a variety of physiological processes including tissue homeostasis, neurotransmission, and inflammation12, 13. Nitric oxide is the product of NO synthase (NOS), which catalyzes the conversion of arginine and oxygen (O2) into NO and citrulline (see Figure 1). There are three isoforms of NOS, each encoded by a different gene. Two of these isoforms, endothelial NOS (eNOS) and neuronal NOS (nNOS), are expressed constitutively at low levels; these enzymes produce picomolar concentrations of NO. Endothelial NOS and nNOS contain a calmodulin-binding domain and are activated by calmodulin in a calcium-dependent manner. The third isoform, inducible NOS (iNOS), is calcium-independent, as it binds calmodulin with high affinity under calcium replete physiological conditions14. In contrast, iNOS is not expressed under normal conditions, but is induced in high levels during inflammation. Once iNOS is expressed, it produces NO in nanomolar to micromolar concentrations15.
Figure 1. NO Production.

Nitric oxide synthase (NOS) catalyzes the reaction between arginine and oxygen, with the consumption of 1.5 moles of NADPH per mole of NO produced. Citrulline and water are also produced during the reaction.
A thorough knowledge of the chemical properties of NO is important for understanding its biologic effects. Nitric oxide is an intermediate oxidation product of nitrogen; it has an unpaired electron and therefore can be considered a free radical. In the presence of other free radicals, NO propagates free radical induced damage. It reacts readily with oxidants to form more stable, higher nitrogen oxides such as nitrite and nitrate, which accounts for the short life span of NO under physiologic conditions. Furthermore, NO diffuses rapidly away from production sites. Because it is soluble in both water and lipids, NO can efficiently diffuse across biologic membranes and through cells without requiring cell receptors or channels16.
Nitric oxide signaling is complex, and much of its functionality is based on the rapid movement of the molecule. Because it is freely permeable, NO can rapidly move in and out of neighboring cells over the span of a second. It has a limited life span under physiologic conditions (estimated half-life of 1 sec), and acts locally at a distance equivalent to 4–5 cells. Thus, the paracrine effects of NO decrease with distance as a result of this diffusion capacity rather than actual consumption of the molecule. The combination of rapid diffusion and short life span allows NO to synchronously stimulate and activate groups of cells17.
In vivo, NO is scavenged mainly by reacting with hemoglobin in red blood cells. Because of scavenging by hemoglobin, endothelial cells need to produce 100nM NO to achieve 5–10nM concentrations in adjacent targets, such as smooth muscle cells18. The best understood physiological effects of NO are mediated by cyclic GMP (cGMP). Soluble guanylyl cyclase, the enzyme that catalyzes production of cGMP from GTP contains protoporphyrin IX, with iron in the ferrous state, which binds NO with 10,000 fold greater affinity than molecular oxygen19. Binding of NO activates soluble guanylyl cyclase. Production of cGMP leads to activation of cGMP-dependent protein kinases and cytosolic calcium depletion in target cells. It is believed that cGMP mediates most physiologic effects of NO in many cell types.
The relative specificity and safety of NO are evidenced by the widespread presence of this molecule throughout the vasculature and nervous system. Lack of NO toxicity under normal conditions is due to the fact that eNOS and nNOS do not produce high concentrations of NO. Moreover, NO produced by eNOS rapidly diffuses into red blood cells, where it is converted into nitrate upon reaction with oxyhemoglobin. However, NO produced under inflammatory conditions may be toxic for several reasons. First, induction of iNOS during inflammation leads to local concentrations of NO in the micromolar range, as opposed to nanomolar range under normal conditions15. Second, unlike eNOS-derived NO, iNOS-derived NO cannot be rapidly scavenged by red blood cells. Third, in the absence of the reaction with hemoglobin that produces relatively non-toxic nitrate species, NO may react with superoxide anion, which produces the highly toxic oxidation intermediate peroxynitrite20. Like NO, superoxide is produced at inflammatory sites as a consequence of the induction of oxidative enzyme complexes, such as NADPH oxidases and xanthine oxidase.
Superoxide is normally detoxified by superoxide dismutase (SOD); however, NO reacts with superoxide faster than SOD, thus countering the protective effect of SOD and leading to the production of peroxynitrite instead. Because of the diffusion abilities of NO, superoxide and NO do not need to be produced in the same cell in order to form significant amounts of peroxynitrite. With the close proximity of NO and superoxide at inflammation sites, the formation of peroxynitrite is consistent, and is considered a hallmark of the inflammatory process 21.
Although peroxynitrite is not a free radical, it is much more reactive than its parent molecules NO and O220. Its half-life under physiologic conditions is only 10–20 ms, but this time is sufficient to allow diffusion across membranes, crossing one to two cell spans21. Peroxynitrite exerts its effects on proteins through two mechanisms. It can directly oxidize target protein moieties, including thiols, iron/sulfur centers, and zinc fingers22. These reactions are among the fastest known for peroxynitrite23. In addition to its direct effects, peroxynitrite indirectly mediates oxidation of biomolecules by decomposing into highly reactive radicals, which include the hydroxyl radical and nitrogen dioxide (NO2•). Peroxynitrite can nitrate tyrosine residues in proteins, resulting in the formation of nitrotyrosine. Because nitration of tyrosine is an alternative to phosphorylation at key residues, it can affect a protein’s enzymatic activity and interfere with intracellular signaling processes24.
NO and NEC
Current theory regarding the pathogenesis of NEC suggests that the disease is initiated by perinatal insults to the immature intestine, including formula feeding, bacterial colonization of the gut, hypoxia, and hypoperfusion. These insults result in an initial epithelial injury that may cause intestinal inflammation with release of inflammatory mediators, leading to impaired gut barrier function. Luminal bacteria then translocate across the compromised barrier, further propagating the inflammatory cascade and exacerbating epithelial injury. Extensive barrier failure and ensuing intestinal tissue necrosis are the ultimate manifestations of NEC5, 25. We have shown that NO plays an important role in intestinal epithelial injury in NEC. The role of iNOS-derived NO was strongly implicated first by the work of Ford and colleagues, who found an upregulation of iNOS mRNA and protein in infants undergoing laparotomy for NEC, as compared to those undergoing intestinal resection for other reasons11.
Both constitutive and inducible isoforms of NOS are expressed in the gastrointestinal tract. The constitutive calcium-dependent eNOS isoform is expressed in the myenteric plexus, endothelial cells, gastric epithelial cells, and enterocytes26, 27. Within the intestinal epithelium, NO is believed to modulate water and electrolyte transport, as well as mucosal permeability28, 29. Inhibition of NOS leads to an increase in transepithelial passage of 51Cr-EDTA, suggesting a role for NO in maintaining mucosal integrity29. NO also acts to enhance mucosal blood flow and maintain microvascular tone. NO inhibits platelet aggregation and adhesion of leukocytes30. Guinea pigs treated with NOS inhibitors develop ileitis25. These data suggest that low levels of NO produced by eNOS may be required for the maintenance of intestinal homeostasis15.
In contrast, sustained overproduction of NO as a result of iNOS upregulation may have cytopathic effects on the epithelium, leading to intestinal barrier failure. As discussed earlier, peroxynitrite, a product of the reaction of NO with superoxide, is largely responsible for the detrimental effects of NO during inflammation. Whereas iNOS is expressed in the ileum of mice colonized with bacteria31, endotoxin (LPS) stimulation significantly increases its expression 27, 32. iNOS mRNA and protein are also upregulated in the rat intestine after LPS administration, leading to derangement in intestinal barrier function. However, treatment with the iNOS inhibitor aminoguanidine ameliorated the epithelial damage and decreased bacterial translocation across the epithelium. Aminoguanidine also decreased bacterial translocation and prevented epithelial sloughing in mice challenged with high doses of LPS33. The iNOS scavenger, NOX, has similar effects: it abrogates LPS-induced bacterial translocation, decreases the levels of enterocyte apoptosis, and inhibits the formation of 3-nitrotyrosine, the molecular footprint of peroxynitrite 34. Moreover, iNOS knockout mice exhibit lower levels of LPS-dependent bacterial translocation than their wild type littermates 35. Taken together, these data indicate a key role of iNOS-derived NO in the gut barrier failure during experimental inflammation.
Sustained NO production has also been found in human patients suffering from various intestinal inflammatory disorders. Patients with active ulcerative colitis have high levels of iNOS and increased 3-nitrotyrosine immunoreactivity in the colonic epithelium36. Helicobacter pylori-induced gastritis is also associated with increased iNOS expression in antral biopsies performed during active disease, as evidenced by 3-nitrotyrosine immunostaining and enterocyte apoptosis37. High levels of iNOS upregulation have been found in resected intestinal tissue from patients with acute NEC11. Interestingly, the expression of iNOS mRNA and protein co-localized with areas of epithelial injury (enterocyte apoptosis and necrosis) along the crypt-villus axis. Furthermore, immunoreactivity to 3-nitrotyrosine was noted in the epithelium and lamina propria at the sites of enterocyte apoptosis. At the time of stoma closure, when the infants had recovered from acute NEC and the inflammation had subsided, the expression of iNOS mRNA had returned to basal levels11. Collectively, these data suggest a strong correlation between iNOS expression, NO and peroxynitrite production on the one hand, and mucosal injury on the other hand. The induction of intestinal barrier failure through enterocyte apoptosis and failure of appropriate restitution of the epithelium are believed to be key.
To better understand the role of NO in the pathogenesis of NEC, we employed an animal model. Many experimental models of NEC use invasive surgical procedures to induce ischemia, or intravenous infusion of pro-inflammatory mediators to induce inflammation. While these models may reproduce the morphological characteristics of human NEC, they do not parallel the pathophysiology of the disease. Our lab has developed a rat model of NEC whereby newborn rat pups are subjected to formula feeding and hypoxia thrice daily; control animals are breast-fed by their mothers. A high percentage of animals from the formula/hypoxia group develop varying degrees of intestinal inflammation, characterized by submucosal edema, neutrophil infiltration, and epithelial sloughing38. This model relies on perinatal insults similar to those associated with, and reproduces pathology similar to, human NEC. It has also been a valuable tool to study the expression profile of inflammatory mediators and iNOS in experimental NEC. Importantly, the pattern of cytokine expression and iNOS upregulation in this model parallels that seen in human NEC. In the rat model of NEC, there is increased iNOS and IFNγ, with decreased IL-12 production38. This model has recently been adapted to mice39–41, allowing the use of genetically modified animals to address the roles of specific genes in the pathogenesis of NEC. We are currently using iNOS knockout mice to further define the role of iNOS-derived NO in the pathogenesis of NEC.
Effects of NO and peroxynitrite on the intestinal epithelium: Apoptosis, Proliferation and Migration
NO and Apoptosis pathways
The mucosal surface of the gastrointestinal tract consists of a vast epithelial monolayer that serves as a physical barrier to the entry of intraluminal antigens and microorganisms. Maintenance of intestinal barrier integrity is the result of a dynamic equilibrium between tissue injury and tissue repair mechanisms such as epithelial restitution and proliferation. Factors that may affect gut barrier integrity include enterocyte apoptosis (programmed cell death) and necrosis, as well as disruption of tight junctions or gap junctions that normally play a key role in maintaining epithelial monolayer integrity. Intrinsic mechanisms that serve to restore monolayer integrity following epithelial injury include enterocyte proliferation, epithelial restitution via enterocyte migration, and re-establishment of cell contacts. The mechanisms by which NO and its toxic nitrogen derivative adduct, peroxynitrite, promote their cytopathic effects while concurrently disrupting the repair mechanisms will be discussed.
Enterocyte apoptosis at the villus tip is a normal physiologic process that plays a key role in the continuous renewal of the intestinal epithelium. However, we and others have reported increased enterocyte apoptosis in both human and experimental NEC11, 38. The increased enterocyte apoptosis co-localizes with upregulation of iNOS activity as well as 3-nitrotyrosine immunostaining, suggesting that peroxynitrite plays a role in this process. Peroxynitrite has been shown to induce apoptosis in a variety of cell types, including undifferentiated neural cells (PC12), intestinal epithelial cells (T84), HL-60 myeloid cells, macrophages, and neurons42–44. To further delineate the mechanisms by which peroxynitrite exerts its cytopathic effects on the intestinal epithelium, we exposed IEC-6 cells, a non-transformed rat epithelial cell line of crypt origin, to peroxynitrite, and examined the effects on apoptosis, proliferation and migration. Our data show that peroxynitrite induces apoptosis in the IEC-6 cells in a time-and dose-dependent manner, at concentrations that are readily attainable in vivo. Such enterocyte apoptosis involves activation of caspase-345.
NO can induce apoptosis directly or indirectly. NO can reversibly inhibit enzymes that produce free radical intermediates, such as those involved in the mitochondrial electron transport chain (ETC). When NO inhibits cytochrome-c oxidase, superoxide may transiently leak from the ETC, leading to the formation of peroxynitrite46. While the reactions of NO with enzymes are usually reversible, the reactions of peroxynitrite are not. Through oxidation of critical cysteine residues, peroxynitrite irreversibly inactivates many components of the ETC, including complex I, complex II, complex III and complex V47, leading to inhibition of oxidative phosphorylation and ultimately, cell death. Experiments in our laboratory have shown that exposure of IEC-6 cells to peroxynitrite de-energizes the mitochondria as judged by 3,3′ dihexyloxacarbocyanine iodide (Mitosensor) staining (Ford HR, unpublished). Electron microscopy of peroxynitrite-treated cells reveals rounding of mitochondria and disruption of the internal mitochondrial christae48.
Exposure to peroxynitrite leads to cellular damage, ultimately leading to cell death via two possible mechanisms, necrosis or apoptosis. The classification of the type of cell death is based on morphological characteristics, enzymological criteria, functional aspects, or immunological characteristics; however, the distinction between necrosis and apoptosis is not always clear49. Cell death primarily through inhibition of mitochondrial respiration and loss of ATP production is normally classified as necrotic. By contrast, apoptotic cell death is characterized by the ATP-dependent activation of cysteine proteases (caspases). 50. The caspases are a family of cysteine proteases, which may be triggered via the activation of death receptors (extrinsically) or by the permeabilization of the outer membrane of mitochondria (intrinsically).
Although the loss of mitochondrial function is a characteristic of necrosis, mitochondria are intricately involved in the execution of the apoptotic program. Mitochondrial outer membrane permeabilization (MOMP) leads to the release of cytochrome c (cyt c) into the cytosol. Cyt c is a key activator of caspases. Although peroxynitrite may cause necrosis by de-energizing the mitochondria, it can also lead to apoptotic cell death via activation of the caspase machinery. This has been shown in many cell types, including PC1251, fibroblasts52, neurons53, neutrophils54, and endothelial cells55. While the induction of apoptosis in these varied cell types requires different experimental conditions, it appears that MOMP and the resulting efflux of proapoptotic signaling molecules occurs universally following peroxynitrite exposure. Cytosolic cyt c acts in concert with apoptosis activating factor-1 (Apaf-1) to activate pro-caspase 9. Active caspase 9 is an initiator caspase, which in turn activates effector caspases, such as caspases 3 and 7. Active effector caspases execute apoptosis by dismantling key proteins and cellular structures 56. Caspases execute apoptosis not only by activation of pro-apoptotic pathways, but also by mitigating survival signals, such as the mitogen-activated protein kinase (MAPK) and phosphoinositide 3-kinase (PI 3-K) pathways57.
MAPK pathways are another class of targets in peroxynitrite-induced cell death. MAPKs are serine/threonine protein kinases that are activated by mitogens and stresses. MAPK signaling cascades regulate multiple cellular processes including proliferation, differentiation, survival, and apoptosis. There are three families of MAPK: extracellular response kinases (ERK), c-Jun N-terminal kinases (JNK), and p38 stress response kinases.
The ERK pathway is activated by growth factors, as well as oxidants and free radicals58. Peroxynitrite activates ERK in fibroblasts59, neutrophils60 and smooth muscle cells61. In human neutrophils, the activation of ERK by peroxynitrite leads to upregulation and surface expression of β2-integrins, and increased neutrophil adhesion to endothelial cells. ERK activation in these cells also leads to enhanced oxidative burst upon stimulation60. ERK-dependent activation of the oxidative burst may be relevant in NEC, where it may favor conversion of locally produced NO into peroxynitrite. ERK is also believed to play a role in apoptotic signaling, however the existing data are controversial. Depending on the cell type and nature of activating stimulus, ERK may either inhibit or promote apoptosis. Peroxynitrite-induced activation of ERK has been found to promote apoptosis via a p53-dependent mechanism 62–64.
The JNK pathways are activated by environmental stresses. When activated, JNK phosphorylates and thereby activates c-jun, which dimerizes with c-fos to form the transcription factor AP-1. The apoptotic effects of JNK depend on the cell type, inciting stimulus, and local milieu. However, most studies indicate that JNK is pro-apoptotic. Peroxynitrite has been shown to activate JNK in several cell types65–67, and dominant negative JNK may protect cells from apoptosis68. The contribution of JNK pathways into peroxynitrite-induced apoptosis in the intestinal epithelium remains unclear.
The activation of p38 MAPK by peroxynitrite has been linked to apoptosis in a variety of cell lines58. Even at low concentrations, peroxynitrite increases the activating phosphorylation of p38 within minutes of exposure66, 69. The activation of p38 causes upregulation of apoptotic effectors, such as growth arrest and DNA damage (GADD) genes 70. In many of the cell systems studied, pharmacologic inhibition of p38 leads to attenuation of the apoptotic response 63, 70. However, like ERK and JNK, p38 may affect apoptosis in opposite ways. We have shown that p38 upregulates cyclooxygenase-2, a rate-limiting enzyme in the biosynthesis of prostanoids71, leading to increased prostanoid production. Prostanoids play an important role in intestinal epithelial homeostasis and inflammation72, 73 and are known to protect intestinal cells from apoptosis74, 75. Therefore, peroxynitrite-activated p38 may have complex effects on apoptosis, cell survival, and inflammation.
NO in Proliferation and Migration signaling pathways
Proliferation and migration of enterocytes are the two major repair processes in the gut epithelium. Although it is known that proliferation of enterocyte precursors in the crypts is dynamically regulated by the physiologic state of the epithelium, the underlying regulatory mechanisms are not well understood 76. The early observations that NO could modulate critical cell signaling processes through the S-nitrosylation of critical cysteine residues provided a clue to understanding the role of redox regulation in signal transduction77. However, the subsequent discovery of peroxynitrite’s ability to nitrate tyrosine residues suggested that NO might, via formation of peroxynitrite, alter proliferative signaling mediated by protein tyrosine kinases. This shifted attention to phosphorylation cascades, as nitrosylation of these residues leads to modulation of cell signaling processes which rely on tyrosine phosphorylation78. One such proliferation-stimulating tyrosine kinase is epidermal growth factor receptor (EGFR), which is known to play an important role in epithelial restitution.
It has been shown, in a rat model of colitis, that EGF significantly reduces colonic ulceration and inflammation, and also promotes healing of peptic ulcers79. The possibility that peroxynitrite can modulate EGF signaling is strong, as EGF acts through a receptor tyrosine kinase. Binding of EGF to the receptor (EGFR) initiates a tyrosine and serine/threonine phosphorylation cascade, leading to gene transcription, cellular proliferation and differentiation80. In the intestinal cell line Caco-2, exposure to peroxynitrite results in nitration of EGFR, which decreases EGFR tyrosine phosphorylation and proliferation in response to EGF81. The inhibition of the EGFR signaling cascade in vivo would result in decreased enterocyte proliferation and differentiation, which, in combination with the apoptosis induced by peroxynitrite, may contribute to the formation and persistence of bare areas at the villus tips.
Another group of proliferative protein tyrosine kinases is the Src kinase family. This non-receptor kinase family includes Src, Fyn, Yes, Lck, Hck, Fgr, Blk, and Lyn. Src kinases are regulated by phosphorylation of tyrosine residues at positive (Tyr416) and negative (Tyr527) regulatory sites. Src members contain a tyrosine kinase domain and two conserved protein-protein interaction domains, SH2 and SH3, which create docking sites for phosphotyrosine and proline-rich helices 82. Activated Src family members participate in multiple signaling processes; they promote cell survival, stimulate proliferation, and inhibit apoptosis. Downstream effectors include Shc, PI3K, phospholipase C, and focal adhesion kinase (FAK)82. We have demonstrated that exposure of enterocytes to peroxynitrite inhibits serum- and EGF-induced proliferation, decreases Src activating phosphorylation, and increases Src nitration. Moreover, peroxynitrite blunts induction of proliferation by the constitutively active v-Src transgene 83.
It should be noted, however, that peroxynitrite may activate rather than inhibit Src. This has been shown in pancreatic adenocarcinoma cells84, endothelial cells85, and astrocytes86. Opposing effects of peroxynitrite may be due to different steric accessibility of target sites in Src kinase signaling complexes, or differences between Src family members. Two distinct mechanisms of src kinase activation by peroxynitrite have been identified. For example, Hck is activated by peroxynitrite-dependent oxidation of a specific cysteine residue, whereas Lyn can be activated via nitrotyrosine formation at the inhibitory site Tyr52787, 88. In summary, it is possible that modulation of the tyrosine kinase signaling cascades by peroxynitrite decreases enterocyte proliferation, which, in combination with peroxynitrite-induced apoptosis, may contribute to gut barrier failure by the formation of denuded areas of the mucosa.
Enterocyte migration is another potential target of NO in NEC. We have demonstrated that enterocyte migration is inhibited in experimental NEC39, and that enterocyte migration in vitro depends on signaling cascades of the small GTPases Rho, Rac, and Cdc42. Exposure to endotoxin activates RhoA, leading to phospho-FAK expression and the enhancement of focal adhesions. This entrenchment of the enterocyte would effectively prevent cell movement39. Preliminary studies in our lab have shown that Cdc42 is activated during enterocyte migration, using an IEC-6 monolayer wound model. We’ve also found that in enterocytes, Cdc42 acts through PAK (p21 kinase) (Ford, unpublished data). NO affects migration in a Rho-, focal adhesion kinase (FAK)-, and N-terminal src homology (SH2)- containing tyrosine phosphatase-2 (SHP-2)-dependent fashion. Furthermore, NO stimulates phosphorylation of FAK in vivo and in vitro, and inhibits formation of lamellipods by causing accumulation of stress fibers and focal adhesions in vitro39. These findings provide an insight into the contribution of NO to the inhibition of enterocyte migration seen in NEC. Effects of NO/peroxynitrite on enterocyte migration machinery warrant further investigation.
Conclusion
Necrotizing enterocolitis is a complex disease characterized by intestinal barrier failure. Such derangement in intestinal barrier function plays a critical role in the progression and severity of the disease. Once barrier failure occurs, the lamina propria is exposed to increased levels of endotoxin and other bacterial products as a result of bacterial translocation. The net result is activation of the neonatal immune system, which triggers an inflammatory signaling cascade that leads to an exuberant pro-inflammatory response characterized by the release of cytokines, prostanoids, and NO. The reaction of NO with superoxide leads to the production of peroxynitrite, a potent oxidant. These two molecules exert their cytopathic effects through various molecular pathways outlined above, which lead to enterocyte apoptosis or necrosis and impairment of both enterocyte proliferation and epithelial restitution through enterocyte migration. This imbalance between tissue injury and repair propagates the inflammatory cascade leading to full blown NEC. The ultimate consequence of these insults is further epithelial injury, intestinal perforation, and systemic sepsis, which can eventually lead to death in the most severe cases. Novel therapies that are designed to neutralize these pro-inflammatory processes are key areas of research in the continued effort to decrease the incidence and severity of NEC in neonates.
Figure 2. Effects of NO on the intestinal barrier.
Perinatal insults and hypoxia lead to mucosal damage. NO and its reactive intermediary, peroxynitrite (ONOO−), are produced. This leads to a cascade of events, including apoptosis at the villus tip, bacterial translocation, and immunocyte activation. Activated immunocytes and inflamed enterocytes continue to produce cytokines and inflammatory mediators such as NO. Unchecked, this cycle will continue and lead to systemic sepsis, and can eventually lead to death in the most severe cases.
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.
Contributor Information
Nikunj K. Chokshi, Department of Pediatric Surgery, Childrens Hospital Los Angeles, Los Angeles, CA 90027.
Catherine J. Hunter, Department of Pediatric Surgery, Childrens Hospital Los Angeles, Los Angeles, CA 90027.
Yigit S. Guner, Department of Pediatric Surgery, Childrens Hospital Los Angeles, Los Angeles, CA 90027.
Anatoly Grishin, Department of Pediatric Surgery, Childrens Hospital Los Angeles, Los Angeles, CA 90027.
Henri R. Ford, Childrens Hospital Los Angeles, Professor and Vice-Chairman, Department of Surgery, Keck School of Medicine, University of Southern California.
References
- 1.Holman RC, Stoll BJ, Curns AT, Yorita KL, Steiner CA, Schonberger LB. Necrotising enterocolitis hospitalisations among neonates in the United States. Paediatr Perinat Epidemiol. 2006 Nov;20(6):498–506. doi: 10.1111/j.1365-3016.2006.00756.x. [DOI] [PubMed] [Google Scholar]
- 2.Rowe MI, Reblock KK, Kurkchubasche AG, Healey PJ. Necrotizing enterocolitis in the extremely low birth weight infant. J Pediatr Surg. 1994 Aug;29(8):987–990. doi: 10.1016/0022-3468(94)90264-x. discussion 990–981. [DOI] [PubMed] [Google Scholar]
- 3.Holman RC, Stoll BJ, Clarke MJ, Glass RI. The epidemiology of necrotizing enterocolitis infant mortality in the United States. Am J Public Health. 1997 Dec;87(12):2026–2031. doi: 10.2105/ajph.87.12.2026. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Llanos AR, Moss ME, Pinzon MC, Dye T, Sinkin RA, Kendig JW. Epidemiology of neonatal necrotising enterocolitis: a population-based study. Paediatr Perinat Epidemiol. 2002 Oct;16(4):342–349. doi: 10.1046/j.1365-3016.2002.00445.x. [DOI] [PubMed] [Google Scholar]
- 5.Hackam DJ, Upperman JS, Grishin A, Ford HR. Disordered enterocyte signaling and intestinal barrier dysfunction in the pathogenesis of necrotizing enterocolitis. Semin Pediatr Surg. 2005 Feb;14(1):49–57. doi: 10.1053/j.sempedsurg.2004.10.025. [DOI] [PubMed] [Google Scholar]
- 6.Sharma R, Tepas JJ, 3rd, Hudak ML, et al. Neonatal gut barrier and multiple organ failure: role of endotoxin and proinflammatory cytokines in sepsis and necrotizing enterocolitis. J Pediatr Surg. 2007 Mar;42(3):454–461. doi: 10.1016/j.jpedsurg.2006.10.038. [DOI] [PubMed] [Google Scholar]
- 7.DeMeo MT, Mutlu EA, Keshavarzian A, Tobin MC. Intestinal permeation and gastrointestinal disease. J Clin Gastroenterol. 2002 Apr;34(4):385–396. doi: 10.1097/00004836-200204000-00003. [DOI] [PubMed] [Google Scholar]
- 8.Neutra MR, Mantis NJ, Kraehenbuhl JP. Collaboration of epithelial cells with organized mucosal lymphoid tissues. Nat Immunol. 2001 Nov;2(11):1004–1009. doi: 10.1038/ni1101-1004. [DOI] [PubMed] [Google Scholar]
- 9.Caplan MS, Sun XM, Hseuh W, Hageman JR. Role of platelet activating factor and tumor necrosis factor-alpha in neonatal necrotizing enterocolitis. J Pediatr. 1990 Jun;116(6):960–964. doi: 10.1016/s0022-3476(05)80661-4. [DOI] [PubMed] [Google Scholar]
- 10.Ford H, Reblock K, Teramana C, et al. The role of nitric oxide in the pathogenesis of necrotizing enterocolitis. Surg Forum. 1995;46:642–644. [Google Scholar]
- 11.Ford H, Watkins S, Reblock K, Rowe M. The role of inflammatory cytokines and nitric oxide in the pathogenesis of necrotizing enterocolitis. Journal of Pediatric Surgery. 1997;32(2):275–282. doi: 10.1016/s0022-3468(97)90194-9. [DOI] [PubMed] [Google Scholar]
- 12.Chang K, Lee SJ, Cheong I, et al. Nitric oxide suppresses inducible nitric oxide synthase expression by inhibiting post-translational modification of IkappaB. Exp Mol Med. 2004 Aug 31;36(4):311–324. doi: 10.1038/emm.2004.42. [DOI] [PubMed] [Google Scholar]
- 13.Yu Z, Xia X, Kone BC. Expression profile of a human inducible nitric oxide synthase promoter reporter in transgenic mice during endotoxemia. Am J Physiol Renal Physiol. 2005 Jan;288(1):F214–220. doi: 10.1152/ajprenal.00258.2004. [DOI] [PubMed] [Google Scholar]
- 14.Spratt DE, Taiakina V, Palmer M, Guillemette JG. Differential binding of calmodulin domains to constitutive and inducible nitric oxide synthase enzymes. Biochemistry. 2007 Jul 17;46(28):8288–8300. doi: 10.1021/bi062130b. [DOI] [PubMed] [Google Scholar]
- 15.Nathan C, Xie Q. Nitric oxide synthases: roles, tolls, and controls. Cell. 1994;78:915–918. doi: 10.1016/0092-8674(94)90266-6. [DOI] [PubMed] [Google Scholar]
- 16.Pacher P, Beckman JS, Liaudet L. Nitric oxide and peroxynitrite in health and disease. Physiol Rev. 2007 Jan;87(1):315–424. doi: 10.1152/physrev.00029.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Lancaster JR., Jr Simulation of the diffusion and reaction of endogenously produced nitric oxide. Proc Natl Acad Sci U S A. 1994 Aug 16;91(17):8137–8141. doi: 10.1073/pnas.91.17.8137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Tsoukias NM, Kavdia M, Popel AS. A theoretical model of nitric oxide transport in arterioles: frequency- vs. amplitude-dependent control of cGMP formation. Am J Physiol Heart Circ Physiol. 2004 Mar;286(3):H1043–1056. doi: 10.1152/ajpheart.00525.2003. [DOI] [PubMed] [Google Scholar]
- 19.Sharma VS, Traylor TG, Gardiner R, Mizukami H. Reaction of nitric oxide with heme proteins and model compounds of hemoglobin. Biochemistry. 1987 Jun 30;26(13):3837–3843. doi: 10.1021/bi00387a015. [DOI] [PubMed] [Google Scholar]
- 20.Beckman JS. Ischaemic injury mediator. Nature. 1990 May 3;345(6270):27–28. doi: 10.1038/345027b0. [DOI] [PubMed] [Google Scholar]
- 21.Denicola A, Souza JM, Radi R. Diffusion of peroxynitrite across erythrocyte membranes. Proc Natl Acad Sci U S A. 1998 Mar 31;95(7):3566–3571. doi: 10.1073/pnas.95.7.3566. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Beckman JS, Koppenol WH. Nitric oxide, superoxide, and peroxynitrite: the good, the bad, and ugly. Am J Physiol. 1996;271(5 Pt 1):C1424–1437. doi: 10.1152/ajpcell.1996.271.5.C1424. [DOI] [PubMed] [Google Scholar]
- 23.Alvarez B, Radi R. Peroxynitrite reactivity with amino acids and proteins. Amino Acids. 2003 Dec;25(3–4):295–311. doi: 10.1007/s00726-003-0018-8. [DOI] [PubMed] [Google Scholar]
- 24.Schopfer FJ, Baker PR, Freeman BA. NO-dependent protein nitration: a cell signaling event or an oxidative inflammatory response? Trends Biochem Sci. 2003 Dec;28(12):646–654. doi: 10.1016/j.tibs.2003.10.006. [DOI] [PubMed] [Google Scholar]
- 25.Upperman JS, Potoka D, Grishin A, Hackam D, Zamora R, Ford HR. Mechanisms of nitric oxide-mediated intestinal barrier failure in necrotizing enterocolitis. Semin Pediatr Surg. 2005 Aug;14(3):159–166. doi: 10.1053/j.sempedsurg.2005.05.004. [DOI] [PubMed] [Google Scholar]
- 26.Nowicki PT, Caniano DA, Hammond S, et al. Endothelial nitric oxide synthase in human intestine resected for necrotizing enterocolitis. J Pediatr. 2007 Jan;150(1):40–45. doi: 10.1016/j.jpeds.2006.09.029. [DOI] [PubMed] [Google Scholar]
- 27.Tepperman BL, Brown JF, Whittle BJ. Nitric oxide synthase induction and intestinal epithelial cell viability in rats. Am J Physiol. 1993;265(2 Pt 1):G214–218. doi: 10.1152/ajpgi.1993.265.2.G214. [DOI] [PubMed] [Google Scholar]
- 28.Barry M, Aloisi J, Pickering S, Yeo C. Nitric oxide modulates water and electrolyte transport in the ileum. Ann Surg. 1994;219:382–388. doi: 10.1097/00000658-199404000-00009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Kubes P. Nitric oxide modulates epithelial permeability in the feline small intestine. Am J Physiol. 1992;262:G1138–G1142. doi: 10.1152/ajpgi.1992.262.6.G1138. [DOI] [PubMed] [Google Scholar]
- 30.Hutcheson IR, Whittle BJ, Boughton-Smith NK. Role of nitric oxide in maintaining vascular integrity in endotoxin-induced acute intestinal damage in the rat. Br J Pharmacol. 1990;101(4):815–820. doi: 10.1111/j.1476-5381.1990.tb14163.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Hoffman RA, Zhang G, Nussler NC, et al. Constitutive expression of inducible nitric oxide synthase in the mouse ileal mucosa. Am J Physiol. 1997;272(2 Pt 1):G383–392. doi: 10.1152/ajpgi.1997.272.2.G383. [DOI] [PubMed] [Google Scholar]
- 32.Chen K, Inoue M, Okada A. Expression of inducible nitric oxide synthase mRNA in rat digestive tissues after endotoxin and its role in intestinal mucosal injury. Biochem Biophys Res Commun. 1996;224(3):703–708. doi: 10.1006/bbrc.1996.1087. [DOI] [PubMed] [Google Scholar]
- 33.Sorrells DL, Friend C, Koltuksuz U, et al. Inhibition of nitric oxide with aminoguanidine reduces bacterial translocation after endotoxin challenge in vivo. Arch Surg. 1996;131(11):1155–1163. doi: 10.1001/archsurg.1996.01430230037007. [DOI] [PubMed] [Google Scholar]
- 34.Dickinson E, Tuncer R, Nadler E, et al. NOX, A nitric oxide scavenger reduces bacterial translocation in rats after endotoxemia by decreasing enterocyte apoptosis. Am J Physiol. 1999;277:G1281–G1287. doi: 10.1152/ajpgi.1999.277.6.G1281. [DOI] [PubMed] [Google Scholar]
- 35.Mishima S, Xu D, Lu Q, Deitch EA. Bacterial translocation is inhibited in inducible nitric oxide synthase knockout mice after endotoxin challenge but not in a model of bacterial overgrowth. Arch Surg. 1997;132(11):1190–1195. doi: 10.1001/archsurg.1997.01430350040007. [DOI] [PubMed] [Google Scholar]
- 36.Singer II, Kawka DW, Scott S, et al. Expression of inducible nitric oxide synthase and nitrotyrosine in colonic epithelium in inflammatory bowel disease. Gastroenterology. 1996;111(4):871–885. doi: 10.1016/s0016-5085(96)70055-0. [DOI] [PubMed] [Google Scholar]
- 37.Mannick EE, Bravo LE, Zarama G, et al. Inducible nitric oxide synthase, nitrotyrosine, and apoptosis in Helicobacter pylori gastritis: effect of antibiotics and antioxidants. Cancer Res. 1996;56(14):3238–3243. [PubMed] [Google Scholar]
- 38.Nadler EP, Dickinson E, Knisely A, et al. Expression of inducible nitric oxide synthase and interleukin-12 in experimental necrotizing enterocolitis. J Surg Res. 2000;92(1):71–77. doi: 10.1006/jsre.2000.5877. [DOI] [PubMed] [Google Scholar]
- 39.Cetin S, Leaphart CL, Li J, et al. Nitric oxide inhibits enterocyte migration through activation of RhoA-GTPase in a SHP-2-dependent manner. Am J Physiol Gastrointest Liver Physiol. 2007 May;292(5):G1347–1358. doi: 10.1152/ajpgi.00375.2006. [DOI] [PubMed] [Google Scholar]
- 40.Jilling T, Simon D, Lu J, et al. The roles of bacteria and TLR4 in rat and murine models of necrotizing enterocolitis. J Immunol. 2006 Sep 1;177(5):3273–3282. doi: 10.4049/jimmunol.177.5.3273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Halpern MD, Khailova L, Molla Hosseini D, et al. Decreased Development of Necrotizing Enterocolitis in IL-18 Deficient Mice. Am J Physiol Gastrointest Liver Physiol. 2007 Oct 18; doi: 10.1152/ajpgi.00168.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Estevez AG, Radi R, Barbeito L, Shin JT, Thompson JA, Beckman JS. Peroxynitrite-induced cytotoxicity in PC12 cells: evidence for an apoptotic mechanism differentially modulated by neurotrophic factors. J Neurochem. 1995;65(4):1543–1550. doi: 10.1046/j.1471-4159.1995.65041543.x. [DOI] [PubMed] [Google Scholar]
- 43.Lin KT, Xue JY, Nomen M, Spur B, Wong PY. Peroxynitrite-induced apoptosis in HL-60 cells. J Biol Chem. 1995;270(28):16487–16490. doi: 10.1074/jbc.270.28.16487. [DOI] [PubMed] [Google Scholar]
- 44.Sandoval M, Zhang XJ, Liu X, Mannick EE, Clark DA, Miller MJ. Peroxynitrite-induced apoptosis in T84 and RAW 264.7 cells: attenuation by L-ascorbic acid. Free Radic Biol Med. 1997;22(3):489–495. doi: 10.1016/s0891-5849(96)00374-7. [DOI] [PubMed] [Google Scholar]
- 45.Potoka DA, Nadler EP, Zhou X, Zhang XR, Upperman JS, Ford HR. Inhibition of NF-kappaB by IkappaB prevents cytokine-induced NO production and promotes enterocyte apoptosis in vitro. Shock. 2000 Sep;14(3):366–373. doi: 10.1097/00024382-200014030-00022. [DOI] [PubMed] [Google Scholar]
- 46.Shiva S, Brookes PS, Patel RP, Anderson PG, Darley-Usmar VM. Nitric oxide partitioning into mitochondrial membranes and the control of respiration at cytochrome c oxidase. Proc Natl Acad Sci U S A. 2001 Jun 19;98(13):7212–7217. doi: 10.1073/pnas.131128898. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Radi R, Cassina A, Hodara R. Nitric oxide and peroxynitrite interactions with mitochondria. Biol Chem. 2002 Mar–Apr;383(3–4):401–409. doi: 10.1515/BC.2002.044. [DOI] [PubMed] [Google Scholar]
- 48.Upperman JS, Nadler EP, Hoffman R, et al. Peroxynitrite (ONOO) induces enterocyte apoptosis in vitro by modulating mitochondrial membrane potential and pars cleavage. Paper presented at: American College of Surgeons Clinical Congress; San Francisco, CA. 1999. [Google Scholar]
- 49.Galluzzi L, Maiuri MC, Vitale I, et al. Cell death modalities: classification and pathophysiological implications. Cell Death Differ. 2007 Jul;14(7):1237–1243. doi: 10.1038/sj.cdd.4402148. [DOI] [PubMed] [Google Scholar]
- 50.Leist M, Single B, Castoldi AF, Kuhnle S, Nicotera P. Intracellular adenosine triphosphate (ATP) concentration: a switch in the decision between apoptosis and necrosis. J Exp Med. 1997 Apr 21;185(8):1481–1486. doi: 10.1084/jem.185.8.1481. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Spear N, Estevez AG, Johnson GV, Bredesen DE, Thompson JA, Beckman JS. Enhancement of peroxynitrite-induced apoptosis in PC12 cells by fibroblast growth factor-1 and nerve growth factor requires p21Ras activation and is suppressed by Bcl-2. Arch Biochem Biophys. 1998 Aug 1;356(1):41–45. doi: 10.1006/abbi.1998.0741. [DOI] [PubMed] [Google Scholar]
- 52.Raghuram N, Fortenberry JD, Owens ML, Brown LA. Effects of exogenous nitric oxide and hyperoxia on lung fibroblast viability and DNA fragmentation. Biochem Biophys Res Commun. 1999 Sep 7;262(3):685–691. doi: 10.1006/bbrc.1999.1216. [DOI] [PubMed] [Google Scholar]
- 53.Estevez AG, Spear N, Manuel SM, Barbeito L, Radi R, Beckman JS. Role of endogenous nitric oxide and peroxynitrite formation in the survival and death of motor neurons in culture. Prog Brain Res. 1998;118:269–280. doi: 10.1016/s0079-6123(08)63214-8. [DOI] [PubMed] [Google Scholar]
- 54.Fortenberry JD, Owens ML, Brown MR, Atkinson D, Brown LA. Exogenous nitric oxide enhances neutrophil cell death and DNA fragmentation. Am J Respir Cell Mol Biol. 1998 Mar;18(3):421–428. doi: 10.1165/ajrcmb.18.3.2875. [DOI] [PubMed] [Google Scholar]
- 55.Dickhout JG, Hossain GS, Pozza LM, Zhou J, Lhotak S, Austin RC. Peroxynitrite causes endoplasmic reticulum stress and apoptosis in human vascular endothelium: implications in atherogenesis. Arterioscler Thromb Vasc Biol. 2005 Dec;25(12):2623–2629. doi: 10.1161/01.ATV.0000189159.96900.d9. [DOI] [PubMed] [Google Scholar]
- 56.Graf D, Bode JG, Haussinger D. Caspases and receptor cleavage. Arch Biochem Biophys. 2007 Jun 15;462(2):162–170. doi: 10.1016/j.abb.2007.03.029. [DOI] [PubMed] [Google Scholar]
- 57.Widmann C, Gibson S, Johnson GL. Caspase-dependent cleavage of signaling proteins during apoptosis. A turn-off mechanism for anti-apoptotic signals. J Biol Chem. 1998 Mar 20;273(12):7141–7147. doi: 10.1074/jbc.273.12.7141. [DOI] [PubMed] [Google Scholar]
- 58.Martindale JL, Holbrook NJ. Cellular response to oxidative stress: signaling for suicide and survival. J Cell Physiol. 2002 Jul;192(1):1–15. doi: 10.1002/jcp.10119. [DOI] [PubMed] [Google Scholar]
- 59.Zhang P, Wang YZ, Kagan E, Bonner JC. Peroxynitrite targets the epidermal growth factor receptor, Raf-1, and MEK independently to activate MAPK. J Biol Chem. 2000 Jul 21;275(29):22479–22486. doi: 10.1074/jbc.M910425199. [DOI] [PubMed] [Google Scholar]
- 60.Zouki C, Zhang SL, Chan JS, Filep JG. Peroxynitrite induces integrin-dependent adhesion of human neutrophils to endothelial cells via activation of the Raf-1/MEK/Erk pathway. Faseb J. 2001 Jan;15(1):25–27. doi: 10.1096/fj.00-0521fje. [DOI] [PubMed] [Google Scholar]
- 61.Upmacis RK, Deeb RS, Resnick MJ, et al. Involvement of the mitogen-activated protein kinase cascade in peroxynitrite-mediated arachidonic acid release in vascular smooth muscle cells. Am J Physiol Cell Physiol. 2004 Jun;286(6):C1271–1280. doi: 10.1152/ajpcell.00143.2003. [DOI] [PubMed] [Google Scholar]
- 62.Yoo K, Choi JW, Choi MS, et al. Mitogen-activated protein kinases (MAPKs) mediate SIN-1/glucose deprivation-induced death in rat primary astrocytes. Arch Pharm Res. 2005 Aug;28(8):942–947. doi: 10.1007/BF02973881. [DOI] [PubMed] [Google Scholar]
- 63.Nabeyrat E, Jones GE, Fenwick PS, Barnes PJ, Donnelly LE. Mitogen-activated protein kinases mediate peroxynitrite-induced cell death in human bronchial epithelial cells. Am J Physiol Lung Cell Mol Physiol. 2003 Jun;284(6):L1112–1120. doi: 10.1152/ajplung.00178.2002. [DOI] [PubMed] [Google Scholar]
- 64.Kaji T, Kaieda I, Hisatsune T, Kaminogawa S. 3-Morpholinosydnonimine hydrochloride induces p53-dependent apoptosis in murine primary neural cells: a critical role for p21(ras)-MAPK-p19(ARF) pathway. Nitric Oxide. 2002 Mar;6(2):125–134. doi: 10.1006/niox.2001.0389. [DOI] [PubMed] [Google Scholar]
- 65.Andreone TL, O’Connor M, Denenberg A, Hake PW, Zingarelli B. Poly(ADP-ribose) polymerase-1 regulates activation of activator protein-1 in murine fibroblasts. J Immunol. 2003 Feb 15;170(4):2113–2120. doi: 10.4049/jimmunol.170.4.2113. [DOI] [PubMed] [Google Scholar]
- 66.Go YM, Patel RP, Maland MC, et al. Evidence for peroxynitrite as a signaling molecule in flow-dependent activation of c-Jun NH(2)-terminal kinase. Am J Physiol. 1999 Oct;277(4 Pt 2):H1647–1653. doi: 10.1152/ajpheart.1999.277.4.H1647. [DOI] [PubMed] [Google Scholar]
- 67.Schieke SM, Briviba K, Klotz LO, Sies H. Activation pattern of mitogen-activated protein kinases elicited by peroxynitrite: attenuation by selenite supplementation. FEBS Lett. 1999 Apr 9;448(2–3):301–303. doi: 10.1016/s0014-5793(99)00372-5. [DOI] [PubMed] [Google Scholar]
- 68.Shrivastava P, Pantano C, Watkin R, et al. Reactive nitrogen species-induced cell death requires Fas-dependent activation of c-Jun N-terminal kinase. Mol Cell Biol. 2004 Aug;24(15):6763–6772. doi: 10.1128/MCB.24.15.6763-6772.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Eligini S, Habib A, Lebret M, Creminon C, Levy-Toledano S, Maclouf J. Induction of cyclo-oxygenase-2 in human endothelial cells by SIN-1 in the absence of prostaglandin production. Br J Pharmacol. 2001 Aug;133(7):1163–1171. doi: 10.1038/sj.bjp.0704163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Oh-Hashi K, Maruyama W, Isobe K. Peroxynitrite induces GADD34, 45, and 153 VIA p38 MAPK in human neuroblastoma SH-SY5Y cells. Free Radic Biol Med. 2001 Jan 15;30(2):213–221. doi: 10.1016/s0891-5849(00)00461-5. [DOI] [PubMed] [Google Scholar]
- 71.Grishin AV, Wang J, Potoka DA, et al. Lipopolysaccharide induces cyclooxygenase-2 in intestinal epithelium via a noncanonical p38 MAPK pathway. J Immunol. 2006 Jan 1;176(1):580–588. doi: 10.4049/jimmunol.176.1.580. [DOI] [PubMed] [Google Scholar]
- 72.Lugo B, Ford HR, Grishin A. Molecular signaling in necrotizing enterocolitis: regulation of intestinal COX-2 expression. J Pediatr Surg. 2007 Jul;42(7):1165–1171. doi: 10.1016/j.jpedsurg.2007.02.006. [DOI] [PubMed] [Google Scholar]
- 73.Chung DH, Ethridge RT, Kim S, et al. Molecular mechanisms contributing to necrotizing enterocolitis. Ann Surg. 2001 Jun;233(6):835–842. doi: 10.1097/00000658-200106000-00014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Hoshino T, Tsutsumi S, Tomisato W, Hwang HJ, Tsuchiya T, Mizushima T. Prostaglandin E2 protects gastric mucosal cells from apoptosis via EP2 and EP4 receptor activation. J Biol Chem. 2003 Apr 11;278(15):12752–12758. doi: 10.1074/jbc.M212097200. [DOI] [PubMed] [Google Scholar]
- 75.Joseph RR, Yazer E, Hanakawa Y, Stadnyk AW. Prostaglandins and activation of AC/cAMP prevents anoikis in IEC-18. Apoptosis. 2005 Dec;10(6):1221–1233. doi: 10.1007/s10495-005-2049-y. [DOI] [PubMed] [Google Scholar]
- 76.Podolsky DK. Regulation of intestinal epithelial proliferation: a few answers, many questions. Am J Physiol. 1993 Feb;264(2 Pt 1):G179–186. doi: 10.1152/ajpgi.1993.264.2.G179. [DOI] [PubMed] [Google Scholar]
- 77.Stamler JS, Lamas S, Fang FC. Nitrosylation. the prototypic redox-based signaling mechanism. Cell. 2001 Sep 21;106(6):675–683. doi: 10.1016/s0092-8674(01)00495-0. [DOI] [PubMed] [Google Scholar]
- 78.Gow AJ, Duran D, Malcolm S, Ischiropoulos H. Effects of peroxynitrite-induced protein modifications on tyrosine phosphorylation and degradation. FEBS Lett. 1996 Apr 29;385(1–2):63–66. doi: 10.1016/0014-5793(96)00347-x. [DOI] [PubMed] [Google Scholar]
- 79.Tarnawski AS, Jones MK. The role of epidermal growth factor (EGF) and its receptor in mucosal protection, adaptation to injury, and ulcer healing: involvement of EGF-R signal transduction pathways. J Clin Gastroenterol. 1998;27(Suppl 1):S12–20. doi: 10.1097/00004836-199800001-00004. [DOI] [PubMed] [Google Scholar]
- 80.van der Geer P, Hunter T, Lindberg RA. Receptor protein-tyrosine kinases and their signal transduction pathways. Annu Rev Cell Biol. 1994;10:251–337. doi: 10.1146/annurev.cb.10.110194.001343. [DOI] [PubMed] [Google Scholar]
- 81.Uc A, Kooy NW, Conklin JL, Bishop WP. Peroxynitrite inhibits epidermal growth factor receptor signaling in Caco-2 cells. Dig Dis Sci. 2003 Dec;48(12):2353–2359. doi: 10.1023/b:ddas.0000007874.20403.1d. [DOI] [PubMed] [Google Scholar]
- 82.Roskoski R., Jr Src kinase regulation by phosphorylation and dephosphorylation. Biochem Biophys Res Commun. 2005 May 27;331(1):1–14. doi: 10.1016/j.bbrc.2005.03.012. [DOI] [PubMed] [Google Scholar]
- 83.Potoka DA, Upperman JS, Zhang XR, et al. Peroxynitrite inhibits enterocyte proliferation and modulates Src kinase activity in vitro. Am J Physiol Gastrointest Liver Physiol. 2003 Nov;285(5):G861–869. doi: 10.1152/ajpgi.00412.2002. [DOI] [PubMed] [Google Scholar]
- 84.MacMillan-Crow LA, Greendorfer JS, Vickers SM, Thompson JA. Tyrosine nitration of c-SRC tyrosine kinase in human pancreatic ductal adenocarcinoma. Arch Biochem Biophys. 2000 May 15;377(2):350–356. doi: 10.1006/abbi.2000.1799. [DOI] [PubMed] [Google Scholar]
- 85.Zou MH, Hou XY, Shi CM, et al. Activation of 5′-AMP-activated kinase is mediated through c-Src and phosphoinositide 3-kinase activity during hypoxia-reoxygenation of bovine aortic endothelial cells. Role of peroxynitrite. J Biol Chem. 2003 Sep 5;278(36):34003–34010. doi: 10.1074/jbc.M300215200. [DOI] [PubMed] [Google Scholar]
- 86.Haskew RE, Mongin AA, Kimelberg HK. Peroxynitrite enhances astrocytic volume-sensitive excitatory amino acid release via a src tyrosine kinase-dependent mechanism. J Neurochem. 2002 Aug;82(4):903–912. doi: 10.1046/j.1471-4159.2002.01037.x. [DOI] [PubMed] [Google Scholar]
- 87.Mallozzi C, Di Stasi AM, Minetti M. Nitrotyrosine mimics phosphotyrosine binding to the SH2 domain of the src family tyrosine kinase lyn. FEBS Lett. 2001 Aug 17;503(2–3):189–195. doi: 10.1016/s0014-5793(01)02726-0. [DOI] [PubMed] [Google Scholar]
- 88.Mallozzi C, Di Stasi MA, Minetti M. Peroxynitrite-dependent activation of src tyrosine kinases lyn and hck in erythrocytes is under mechanistically different pathways of redox control. Free Radic Biol Med. 2001 May 15;30(10):1108–1117. doi: 10.1016/s0891-5849(01)00509-3. [DOI] [PubMed] [Google Scholar]

