Articles in PresS. Am J Physiol Lung Cell Mol Physiol (December 30, 2014). doi:10.1152/ajplung.00327.2014

1

Working with Nitric oxide and Hydrogen sulfide in biological systems

2 Shuai Yuan1, Rakesh P. Patel2* and Christopher G Kevil1*

3 4

1

Department of Pathology, Louisiana State University Health Sciences Center, Shreveport, LA

5

2

Department of Pathology, University of Alabama at Birmingham, Birmingham, AL

6 7

*contributed equally to this work

8 9

Corresponding Author

10

Rakesh P. Patel, PhD

11

Department of Pathology

12

BMR2, room 532

13

University of Alabama at Birmingham

14

Birmingham, AL 35226

15

e-mail: [email protected]

16 17 18

1 Copyright © 2014 by the American Physiological Society.

19

Abstract

20 21 22 23 24 25 26 27 28

Nitric oxide (NO) and hydrogen sulfide (H2S) are gasotransmitter molecules important in numerous physiological and pathological processes. While these molecules were first known as environmental toxicants, it is now evident that that they are intricately involved in diverse cellular functions with impact on numerous physiologic and pathogenic processes. NO and H2S share some common characteristics, but also have unique chemical properties which suggest potential complementary interactions between the two in affecting cellular biochemistry and metabolism. Central amongst these is the interactions between NO, H2S and thiols that constitute new ways to regulate protein function, signaling and cellular responses. In this review, we discuss fundamental biochemical principals, molecular functions, measurement methods, and the pathophysiological relevance of NO and H2S.

29 30

Key words: thiol, oxidation, redox, pulmonary, cardiovascular

2

31

Introduction

32 33 34 35 36 37 38 39 40 41 42 43

Historically considered only as industrial pollutants, nitric oxide (NO) and hydrogen sulfide (H2S) are now appreciated as two key physiologically produced signaling molecules which control multiple cellular functions. Appreciation that dysfunction in how these solvated gases affect these functions, either through deficient formation or downstream reactivity has opened up new therapeutic avenues for a host of diseases in all major organ systems including the lung. Our understanding of NO-homeostasis mechanisms are relatively advanced compared to H2S, primarily due to the latter’s discovery as a biologically relevant entity being post-dated by a decade or more compared to NO. While chemically distinct, there are intriguing parallel mechanisms / concepts in NO and H2S biology that include overlapping functions, technical challenges in how each of these gases are measured in biological milieu, appreciation of the mechanisms and factors that modulate how metabolism of each of these is regulated and therapeutic potential for either inhibiting or repleting NO or H2S. In this article, and within the framework outlined above we provide a general overview of NO and H2S biology.

44

Nitric oxide formation: enzymatic and non-enzymatic sources

45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67

Enzymatic sources: Nitric oxide is synthesized by one of three different isoforms of nitric oxide synthase (NOS) that differ in enzymatic activity, how they are regulated (transcriptional, translational and posttranslational mechanisms) and expressed / compartmentalized in tissues as well as the sub-cellular level. These are called NOS I, II and III, corresponding to the inducible (iNOS), neuronal (nNOS) and endothelial (eNOS) isoforms, typically with high, mid and low activities respectively. In the systemic and pulmonary vasculature, eNOS plays a key role in controlling blood flow and maintaining an antithrombotic and anti-inflammatory luminal surface. On the other hand, iNOS is induced by inflammatory stimuli in leukocytes, airway epithelial cells and alveolar macrophages to mediate pathogen killing. However, production of NO at higher concentrations in this setting is also associated with collateral tissue damage mediated by increased oxidation, nitrosation or nitration of biomolecules, which is emblematic of acute lung injury and ARDS. However, this is only a general paradigm. The pleiotropic effects of NO are regulated only in part by high vs low concentrations. It is evident that the cellular source (i.e. compartmentalization) of NO also plays a key role (97, 98, 123). For example iNOS expressed in polymorphonuclear leukocytes (PMNs) regulates sequestration of these cells in the lung vasculature during cecal ligation and puncture (CLP) induced sepsis, but limits PMN infiltration into the alveolar space; on the other hand NO derived from iNOS in lung parenchymal cells plays no role in this process(99). From a biochemical perspective, how iNOS-derived NO produced in PMN’s elicit distinct effects compared to iNOS-derived NO produced in pulmonary endothelial cells is not clear and underscores the importance of compartmentalization with regard to the NO source, and the local concentration of NO where produced. An additional consideration is that selective effects of NO may be mediated by distinct redox derivatives and associated products (e.g. nitrosation and nitrated epitopes), which can have distinct biological activities compared to NO per se. The parameters of concentration, localization and redox conversion are also emerging as critical in the biological actions of H2S.

68 69 70 71

Non-enzymatic sources: More recently, non-enzymatic sources of NO, via reduction of nitrite have emerged as viable NO-production processes in vivo. Widely used as a stable end-product for assessing NO-formation, it is now clear that nitrite is an integral player in NO-homeostasis pathways. Several mechanisms have been identified that operate under physiologic and pathophysiologic conditions that

3

72 73 74 75 76 77

result in the 1-electron reduction of nitrite to NO. Common features of these mechanisms include acceleration of nitrite-reduction as a function of decreasing pH and oxygen tension(30). Intriguingly, in the lung nitrite can elicit signaling responses at low biological (nM-µM) concentrations under normoxia as well. For example, airway epithelial cell repair was stimulated by nitrite via a mechanism that appeared to involve stimulation of oxidative signaling(122). Figure 1 shows a scheme generalizing NO-formation pathways.

78

NO and nitrite therapeutics for airway and pulmonary vascular diseases

79 80 81 82 83 84 85 86 87 88 89 90 91 92 93 94 95 96 97 98 99

Unfortunately, the absence of a selective nitrite scavenger, or a way to selectively decrease nitrite levels in vivo, precludes definitive assessment of endogenous nitrite in NO-homeostasis pathways. However, data from exogenous or therapeutic nitrite clearly demonstrates this possibility. In the lung, inhaled nitrite reversed pulmonary hypertension and was suggested to be more effective compared to inhaled NO. In pre-clinical models, inhaled nitrite did not lead to rebound hypertension once nitrite inhalation was stopped, which is a frequently observed complication associated with acute termination of inhaled NO gas(42). Putative mechanisms of nitrite action involve activation of pulmonary vascular smooth muscle soluble guanylate cyclase (sGC) and the canonical NO-signaling cascade(10). Moreover, nitrite therapy also protects against pulmonary and vascular smooth muscle remodeling after injury and improves lung function after transplantation(4, 88, 111, 139). Additionally, nitrite therapy protects against acute lung injury induced by mechanical ventilation and inhaled chemical toxicants, in the latter of which the risk of developing reactive airways is also reduced(40, 96, 101, 132). Improvement in both inflammatory and permeability components of ALI has been documented, suggesting that nitrite therapy prevents early events in the disease. Interestingly, inhaled NO, which to date is a primary therapeutic for treatment of pulmonary hypertension in the new born, has also been shown to prevent ischemia-reperfusion injury in distal extrapulmonary tissues(26, 64, 65, 84, 117). In this case, the model proposed is that iNO increases the level of circulating nitrite (and perhaps other NO-adducts e.g. S-nitrosothiols) which then stimulate NO-signaling in other tissues. In fact, NO-repletion by nitrite or iNO has proven successful in multiple pre-clinical and clinical studies in different settings of acute and chronic inflammatory tissue injury and supports the model that biological pools of NO-signaling equivalents exist in the form of nitrite in tissues that can be manipulated therapeutically.

100

H2S formation: enzymatic and non-enzymatic pathways

101 102 103 104 105 106 107 108 109 110 111

Enzymatic sources: For nearly a decade, H2S has been discussed as the third biological gasotransmitter along with NO and CO (126). Three distinct H2S generating enzymes have been identified in mammalian systems (Figure 2). The best-characterized H2S producing enzyme, cystathionine β-synthase (CBS), is a pyridoxal-phosphate (PLP) dependent enzyme that catalyzes the condensation between homocysteine and serine, giving rise to cystathionine Owing to the similarity of structures, CBS also uses cysteine, instead of serine, to generate H2S (47). Another PLP dependent enzyme, cysthathionine γ-lyase (CSE), catalyzes the conversion of cystathionine to cysteine, but also utilizes homocysteine and cysteine to generate H2S (16, 138). Additionally, CSE and CBS are crucial hubs in transsulfuration pathways which pass sulfur between various sulfur containing amino acids(50). The third pathway comprises two enzymes found within the mitochondrion. Cysteine aminotransferase (CAT) first converts cysteine to 3-mercaptopyruvate (3-MP), from which H2S is released by 3-mercaptopyruvate sulfurtransferase (3-MST). Importantly, this

4

112 113

process involves two equivalents of reactive thiols or a reduced thioredoxin, with byproducts of a disulfide or an oxidized thioredoxin.

114 115 116 117 118 119 120 121 122 123 124 125 126 127 128 129 130 131 132 133 134

All three enzymes, CBS, CSE and 3-MST, have been documented to be expressed in the lungs, with possible variations between species and cell types (5, 83, 94, 124, 131). The contribution of these enzymes on H2S production is affected by their expression levels and the availability of substrates. For example, CSE is ~60 fold higher than CBS in the mouse liver and accounts >97% H2S production. In comparison, CBS is the predominant H2S producing enzyme in the brain where its expression is higher than CSE (51). Although cysteine and homocysteine are the substrates of both CSE and CBS, CSE is the primary source of H2S in homocysteinemia, while in the kidney where the cysteine level is ~10 fold higher than in the liver or the brain, CBS may predominate. Classically considered as a cytosolic enzyme, CBS is also found to localize in other compartments such as mitochondria(7, 41, 112) and nuclei (1, 52). Its nuclear translocation is related to post-translational modification by small ubiquitin-like modifier (SUMO) which inhibits its activity (49). Moreover, CBS activity is regulated by other small molecules, including its activation by S-adenosylmethionine (SAM) and glutathione (GSH) (46, 86) and inhibition by NO and carbon monoxide (CO) (28, 49, 115). Relatively less is known about regulation of CSE and 3MST. Considered to be the major source of H2S in the cardiovascular system, CSE is expressed in both mouse lung endothelial cells and smooth muscle cells, with its effects on lung-function is poorly explored. 3-MST and CAT are found in most cell types but less abundant than CBS and CSE. Although found in both cytosol and mitochondria (78, 83), their activity may be greater in mitochondria due to higher concentrations of cysteine, the substrate for CAT (49). H2S has been proposed to be an endothelium-derived hyperpolarizing factor (EDHF), based on the deficiency of acetylcholine-induced vasorelaxation in CSE knockout mice (82, 125). However, the proposed mechanism is through KATP sulfhydration via CSE derived H2S, which does not support a direct role of H2S as an EDHF (22).

135 136 137 138 139 140 141 142 143

Non-enzymatic sources: Similar to NO, alternate pathways for H2S formation may exist in the absence of CBS, CSE or 3-MST. An early study showed human erythrocytes reduce elemental sulfur to H2S, with reducing equivalents provided by glutathione, NADH and NADPH(102) (Equation 1). Other forms of sulfane sulfur, where sulfur binding to another sulfur as in thiosulfate (S 2O32-), polysulfide (RSnR) and persulfide (RSSH), may also be reduced to free H2S by low molecular weight reductants independent of enzymatic pathways (49, 70). The yield of H2S via this mechanism may be relatively low but could serve as storage depots of H2S equivalents, as sulfane sulfur can be generated by H2S autoxidation (Equation 2) (70, 73) and by serial oxidation processes in the mitochondria(49). Figure 2 summarizes the transsulfuration pathway and H2S metabolism. 𝐸𝑞𝑢𝑎𝑡𝑖𝑜𝑛 1: 𝑆8 + 16 𝐺𝑆𝐻 → 8 𝐻2 𝑆 + 8 𝐺𝑆𝑆𝐺 𝐺𝑙𝑢𝑡𝑎𝑡ℎ𝑖𝑜𝑛𝑒 𝑅𝑒𝑑𝑢𝑐𝑡𝑎𝑠𝑒

𝐺𝑆𝑆𝐺 + 𝑁𝐴𝐷(𝑃)𝐻 + 𝐻 + →

2 𝐺𝑆𝐻 + 𝑁𝐴𝐷(𝑃)+

𝐸𝑞𝑢𝑎𝑡𝑖𝑜𝑛 2: 𝐻2 𝑆 + 𝑂2 → 𝑆0 + 𝐻2 𝑂2 𝑆0

𝑆0

𝑆0

𝑆0

𝑅𝑆𝐻 → 𝑅𝑆𝑆𝐻 → 𝑅𝑆3 𝐻 → … → 𝑅𝑆𝑛 𝐻 144 145

5

146

NO and H2S Donors

147 148 149 150 151 152 153 154

NO and H2S donors are widely used experimental tools, and potential therapeutics to model and interrogate biological functions. Donors allow introduction of defined amounts with defined rates of release of test compound. For NO, many different NO-donors are commercially available that can expose biological systems to low levels of NO for long periods of time, to high levels of NO for short periods, and variations in between. In addition, NO-donors display pH dependence, sensitivity to oxidants and reductants, which influence what NO-derivatives are produced, and in the most recent generation also offer organelle targeting for NO-release (56). These aspects of NO-donors have been reviewed many times elsewhere. Therefore, we will focus our discussion on H2S donors, which is a rapidly evolving area.

155 156 157 158 159 160

To better understand the pros and cons of available sulfide donors, it is helpful to review the chemical properties of H2S. H2S exists in gaseous form at ambient temperature and pressure. At equilibrium between gas and aqueous phases, its solubility at 37℃ is 80 mM (50). In aqueous solution, H2S dissolves into HS- and S2-, with the pKa of 7 and 12.2-15 respectively at room temperature (87, 133) (Equation 3). Accordingly, it is estimated that two thirds of H2S dissolved in water is HS-, and one third H2S, with a negligible amount of S2- . 𝐸𝑞𝑢𝑎𝑡𝑖𝑜𝑛 3: 𝐻2 𝑆 (𝑔) ⇔

pKa1 =7

𝐻2 𝑆 (𝑎𝑞) ⇔

pKa1 =12−15

𝐻𝑆 − + 𝐻+ ⇔

𝑆 2− + 2𝐻 +

161 162 163 164 165 166 167 168 169

Equation 3 underscores the reversibility between H2S, HS- and S2-. Importantly then in an open system, the equilibrium will be pulled to the left leading to rapid decreases in solvated H2S levels ; note this is also the case with NO. An additional and important consideration is the permeability of H2S into lipid bilayers. While H2S freely diffuses through cell membrane (75), HS- does not. This may affect the availability of H2S in the cytoplasm or other cellular compartments such as mitochondria, although these diffusion limitations may be altered by changes in pH resulting in conversion of HS- to H2S. Therefore, when using sulfide donors, it’s important to consider its cellular accessibility versus its volatility. Currently, available H2S donors fall into three categories: sulfide salts, natural compounds and synthetic compounds.

170 171 172 173 174 175 176 177 178

Sulfide Salts: Sulfide salts, such as Na2S and NaHS, are the simplest H2S generating compounds and most widely used in experimental studies. The release of H2S from Na2S and NaHS is instantaneous upon hydration and dissociation in aqueous solution. H2S release does not reach steady state however due to volatilization of H2S into the gas phase, and thus H2S exposure is transient. The half-life of H2S in plasma is ~20 minutes in humans and ~ 5 minutes in mice(8) after a single bolus injection of Na2S, while the half-life in a cell culture model is also 5 minutes (89). Apart from volatility, H2S is also vulnerable to oxidation. It is recommended that stock solutions be made fresh, sealed in air-tight container with minimal head space, and the concentration confirmed by validated methods. Due to these disadvantages, therapeutic potential of sulfide salts are highly limited.

179 180 181 182 183

Natural Compounds: Garlic consumption has been associated with health benefits for a long time and there are a variety of garlic-based products available and marketed accordingly. Allicin (diallyl thiosulfinate), the best known active ingredient in garlic extracts, decomposes into diallyl disulfide (DADS) and diallyl trisulfide (DATS) (3). These compounds have at least one disulfide bond and an allyl group whose double bonded carbon is adjacent to the sulfur bonded carbon. In the presence of GSH or

6

184 185 186 187 188 189 190 191 192 193

other reduced thiols, a nucleophilic displacement of the α-carbon of this allyl group can generate a hydrodisulfide that is further reduced by GSH to H2S (6, 70). The requirement for cellular metabolism and reducing equivalents is an important consideration in assessing how much H 2S is generated by these compounds and typically allows for slower and more sustained release of H2S compared to sulfide salts. Additionally, these natural products are lipophilic allowing H2S formation to occur in all cell compartments. This hydrophobicity however, is also a limitation as both DADS and DATS are immiscible in water, making delivery of these compounds challenging. Other natural compounds proposed to release H2S include isothiocyanates, such as sulforaphane from broccoli and allyl isothiocyanate, from wasabi (53). However, further understanding of their mechanisms of action and role of H2S in health promoting effects is required.

194 195 196 197 198 199 200 201 202 203 204 205 206 207 208

Synthetic Compounds: Several efforts have been made to develop synthetic H2S donors. GYY4137 is a derivative of Lawesson’s reagent that releases H2S over several minutes via nucleophilic attack and after its dissociation in solution. GYY4137 is believed to release H2S in a prolonged manner, although some studies using amperometry have showed that the release rate was significantly reduced compared to NaHS (69). Characterization of GYY4137 dependent sulfide release in plasma has been reported(67, 69); however, the methylene blue method was used to measure H2S (89). This method for measuring H2S levels is questionable and have we have observed that GYY4137 no longer increased free sulfide 30 minute after application. Considering a high dosage of GYY4137 is required for its physiological effects (400 μmol/L for a plateau concentration of 4h, led to nitrogen dioxide concentrations of 2) to be lower than what we observed. This is likely due to the fact that our chemical work flow for reductant labile sulfide also includes measurement of thiosulfate levels that were not measured using LC-MS methods in the report by Ida et al. However, similar GSSH levels were reported by Lu and colleagues using MBB and LC-MS/MS(72). Together, comparison of these three studies suggests the existence of other reductant-labile sulfur in the tissue, such as protein persulfide/polysulfide and thiosulfate could be biologically relevant forms of sulfide (43, 90).

491 492 493 494 495

14

496 497

Table 2: Levels of Persulfide/Polysulfide and Reductant Labile Sulfur

Tissue Measurement T Ida, et al. (nmol/mg tissue)

GSSH GSSSH CysSSH GSSSG GSSSSG CysSSSSCys CysSSSSSCys Total C Lu, et al. GSSH (nmol/mg protein) X Shen, et al. Reductant Labile Sulfur (nmol/mg protein)

Heart

Lung

0.27 0.02 0.0067 0.0008 0.0195 0.001 0.0005 0.001235 0.00075 0.00015 0.0043 0.0045 0.00525 0.0069 0.307 0.034585

Liver

Brain

0.3 0.002 0.005 0.00375 0.0003 0.006 ND 0.31705

0.77 0.0032 0.011 0.00035 0.00295 ND ND 0.7875

N/A

N/A

~0.2

~0.5

0.4398

1.4508

0.4719

1.392

498 499 500 501 502 503 504 505 506 507 508

Electrochemistry: Similar to NO, H2S can be measured based on its redox properties. The sulfide-specific ion-selective electrodes (ISEs) are well documented but used mostly in non-biological samples. ISEs require an alkaline environment (pH>11) which favor the accumulation of sulfide ion (S2-). In biological samples, however, this could lead to artificially high readings due to hydroxyl replacement of the cysteine sulfur(89). Alternatively, Kraus and colleagues developed a polarographic electrode with a H2Spermeabile membrane, which allows the diffusion of H2S from biological samples at pH 7, into the electrode (pH 10)(21, 60) . The polarographic electrode has the advantage of high sensitivity (10-20nM H2S gas) and the capacity of real-time measurement(129). It may be possible to measurement different sulfide pools by altering sample pH. However, it is sensitive to temperature and pressure and requires frequent calibration.

509 510 511 512 513 514 515 516 517 518 519 520 521

Fluorescent/Chemiluminescent Probes: Sulfide specific fluorescent/chemiluminescent probes are a fastdeveloping tools in H2S research that generally provide the potential for real-time and subcellular measurements. These probes are usually designed according to the reducing and nucleophilic properties of sulfide (76). Current probes provide moderate to good specificity to sulfide compared to other biological thiols. To highlight a few, hydrosulfide naphthalimide-2 (HSN2) has an azido group which can be reduced by H2S into a fluorogenic amine (79). The fluorescence of HSN2 is increased 60 fold by 100 equivalents of H2S. Its specificity to H2S is maintained with the presence of 2,000 equivalents of GSH and cysteine. However, the activation of HSN2 takes 45 minutes and real-time measurement is not possible. Interestingly, Ai and Chen reported a genetically coded azide in GFP, which they used to image endogenous H2S production in HEK 293T cells (15). This theoretically allows sulfide probing at target proteins. Guo and co-workers also developed a ratiometric fluorescent probe, based on nucleophilic addition of H2S to the probe, which is membrane permeable and has little cytotoxicity, therefore allowing real-time H2S probing in living cells (71). Although azide based fluorescent probes usually provide 15

522 523 524 525 526 527

excellent selectivity for H2S over other thiols, excitation with higher power or extended exposure photoactivate these probes. To meet this challenge, Pluth and Bailey developed a chemiluminescent probe, CLSS-2, to minimize noise due to photo-activation (76). It is worth noting that most probes were tested in different conditions (sample concentration, buffer pH, oxygen pressure, etc), which may alter biological sulfide pools. Selectivity and sensitivity of these probes need to be further investigated under physiological conditions and specific cell culture conditions.

528 529 530 531 532 533 534 535 536 537 538 539 540 541

Tag-Switch Technique: Mounting evidence indicates that H2S elicits its biological effects by forming protein persulfides, termed sulfhydration or persulfidation (63, 81, 82, 103). However, due to similar reactivities of persulfides and thiols, it is challenging to identify protein persulfides specifically. The modified biotin switch assay developed by Snyder and colleagues uses methyl methanethiosulfonate (MMTS) to mask free thiols but not persulfide, followed by biotin-HPDP labeling of persulfide (81). However, it is unknown how MMTS spares persulfide specifically. Actually, it has been shown that MMTS may also react with small molecular thiols such as GSH (92). Another method uses iodoacetic acid to label both persulfides and free thiols, after which persulfide adducts are reduced by dithiothreitol (DTT) (63). But the selectivity of DTT reduction also remains unclear. Recently, a thiol blocker, methylsulfonyl benzothiazole (MSBT), developed by Xian et al. was applied in a tag-switch assay (135). MSBT reacts with protein persulfide and yields a highly activated disulfide, which distinguishes it from free thiol adducts and native protein disulfides, so that certain nucleophiles, such as cyanoacetate-biotin, can specifically label persulfides. Additionally, the MSBT based tag-switch technique also showed potential to provide spatial resolution in cell culture (135).

542

Conclusion

543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558

As a gasotransmitter, NO plays fundamental role in vascular function and redox biology. For decades, our understanding of NO biology has evolved through empirical medications, identification of sGC as a receptor that transduces NO-signaling in various pathways, bioavailable metabolites, effects on protein post-translational modifications and novel therapeutic approaches. Although many refined tools have been developed to study NO biology as discussed in this review, not all of them are readily available in each and every laboratory. More importantly, careless choice of tools or misinterpretation of the data often leads to convoluted and false conclusions, especially when various commercial user-friendly kits are employed inappropriately. By comparison, the field of H2S biology is considerably further behind. Our knowledge on H2S chemistry and biology is confusing and contradictory on several aspects. The field is still trying to resolve basic questions such as how are H2S producing enzymes regulated, what are physiological concentrations of H2S, and how does H2S elicit its various biological functions, all of which are limited by available tools and reagents for study. While it is dangerous to ignore flaws in models and methods, especially when identified, it is also unwise to be frightened by difficulty. Without a doubt, the future NO and H2S research will yield interesting and important new discoveries in the years to come. Hopefully this review provides a good introduction to investigators new to the field of NO and H2S that will help them deftly approach study of these molecules in the biological systems.

16

559 560 561 562

Acknowledgements: This research was supported by the CounterACT Program, National Institutes of Health, Office of the Director, and the National Institute of Environmental Health Sciences, Grant Number 1U01ES023759 to RPP and by NIH HL11331 to C.G.K and a Malcolm Feist Cardiovascular Predoctoral Fellowship to S.Y.

563

Conflicts of Interest:

564 565 566

RPP and CGK are co-inventors on patents for use of nitrite salts for the treatment of cardiovascular conditions. C.G.K. has I.P. interest in hydrogen sulfide detection technology and is chief scientific officer and co-founder of Innolyzer LLC.

567 568

17

569 570

References

571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614

1. Agrawal N and Banerjee R. Human polycomb 2 protein is a SUMO E3 ligase and alleviates substrate-induced inhibition of cystathionine beta-synthase sumoylation. PloS one 3: e4032, 2008. 2. Ali MY, Ping CY, Mok YY, Ling L, Whiteman M, Bhatia M, and Moore PK. Regulation of vascular nitric oxide in vitro and in vivo; a new role for endogenous hydrogen sulphide? Br J Pharmacol 149: 625634, 2006. 3. Amagase H. Clarifying the real bioactive constituents of garlic. J Nutr 136: 716S-725S, 2006. 4. Baliga RS, Milsom AB, Ghosh SM, Trinder SL, Macallister RJ, Ahluwalia A, and Hobbs AJ. Dietary nitrate ameliorates pulmonary hypertension: cytoprotective role for endothelial nitric oxide synthase and xanthine oxidoreductase. Circulation 125: 2922-2932, 2012. 5. Baskar R, Li L, and Moore PK. Hydrogen sulfide-induces DNA damage and changes in apoptotic gene expression in human lung fibroblast cells. FASEB journal : official publication of the Federation of American Societies for Experimental Biology 21: 247-255, 2007. 6. Benavides GA, Squadrito GL, Mills RW, Patel HD, Isbell TS, Patel RP, Darley-Usmar VM, Doeller JE, and Kraus DW. Hydrogen sulfide mediates the vasoactivity of garlic. Proc Natl Acad Sci U S A 104: 17977-17982, 2007. 7. Bhattacharyya S, Saha S, Giri K, Lanza IR, Nair KS, Jennings NB, Rodriguez-Aguayo C, LopezBerestein G, Basal E, Weaver AL, Visscher DW, Cliby W, Sood AK, Bhattacharya R, and Mukherjee P. Cystathionine beta-synthase (CBS) contributes to advanced ovarian cancer progression and drug resistance. PloS one 8: e79167, 2013. 8. Bir SC, Kolluru GK, McCarthy P, Shen X, Pardue S, Pattillo CB, and Kevil CG. Hydrogen sulfide stimulates ischemic vascular remodeling through nitric oxide synthase and nitrite reduction activity regulating hypoxia-inducible factor-1alpha and vascular endothelial growth factor-dependent angiogenesis. Journal of the American Heart Association 1: e004093, 2012. 9. Blackstone E, Morrison M, and Roth MB. H2S induces a suspended animation-like state in mice. Science 308: 518, 2005. 10. Blood AB, Schroeder HJ, Terry MH, Merrill-Henry J, Bragg SL, Vrancken K, Liu T, Herring JL, Sowers LC, Wilson SM, and Power GG. Inhaled nitrite reverses hemolysis-induced pulmonary vasoconstriction in newborn lambs without blood participation. Circulation 123: 605-612, 2011. 11. Bruce King S. Potential biological chemistry of hydrogen sulfide (H2S) with the nitrogen oxides. Free radical biology & medicine 55: 1-7, 2013. 12. Bryan NS and Grisham MB. Methods to detect nitric oxide and its metabolites in biological samples. Free radical biology & medicine 43: 645-657, 2007. 13. Butler AR, Calsy-Harrison AM, Glidewell C, and Sørensen PE. The pentacyanonitrosylferrate ion—V. The course of the reactions of nitroprusside with a range of thiols. Polyhedron 7: 1197-1202, 1988. 14. Carballal S, Trujillo M, Cuevasanta E, Bartesaghi S, Moller MN, Folkes LK, Garcia-Bereguiain MA, Gutierrez-Merino C, Wardman P, Denicola A, Radi R, and Alvarez B. Reactivity of hydrogen sulfide with peroxynitrite and other oxidants of biological interest. Free radical biology & medicine 50: 196-205, 2011. 15. Chen ZJ and Ai HW. A highly responsive and selective fluorescent probe for imaging physiological hydrogen sulfide. Biochemistry 53: 5966-5974, 2014. 16. Chiku T, Padovani D, Zhu W, Singh S, Vitvitsky V, and Banerjee R. H2S biogenesis by human cystathionine gamma-lyase leads to the novel sulfur metabolites lanthionine and homolanthionine and is responsive to the grade of hyperhomocysteinemia. J Biol Chem 284: 11601-11612, 2009.

18

615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661

17. Chouchani ET, Methner C, Nadtochiy SM, Logan A, Pell VR, Ding S, James AM, Cocheme HM, Reinhold J, Lilley KS, Partridge L, Fearnley IM, Robinson AJ, Hartley RC, Smith RA, Krieg T, Brookes PS, and Murphy MP. Cardioprotection by S-nitrosation of a cysteine switch on mitochondrial complex I. Nature medicine 19: 753-759, 2013. 18. Cortese-Krott MM, Fernandez BO, Santos JL, Mergia E, Grman M, Nagy P, Kelm M, Butler A, and Feelisch M. Nitrosopersulfide (SSNO(-)) accounts for sustained NO bioactivity of S-nitrosothiols following reaction with sulfide. Redox biology 2: 234-244, 2014. 19. Diers AR, Keszler A, and Hogg N. Detection of S-nitrosothiols. Biochimica et biophysica acta 1840: 892-900, 2014. 20. Dimitrios T. Analysis of nitrite and nitrate in biological fluids by assays based on the Griess reaction: appraisal of the Griess reaction in the L-arginine/nitric oxide area of research. Journal of chromatography B, Analytical technologies in the biomedical and life sciences, 2007. 21. Doeller JE, Isbell TS, Benavides G, Koenitzer J, Patel H, Patel RP, Lancaster JR, Jr., Darley-Usmar VM, and Kraus DW. Polarographic measurement of hydrogen sulfide production and consumption by mammalian tissues. Anal Biochem 341: 40-51, 2005. 22. Edwards G, Feletou M, and Weston AH. Hydrogen sulfide as an endothelium-derived hyperpolarizing factor in rodent mesenteric arteries. Circ Res 110: e13-14; author reply e15-16, 2012. 23. Filipovic MR, Eberhardt M, Prokopovic V, Mijuskovic A, Orescanin-Dusic Z, Reeh P, and Ivanovic-Burmazovic I. Beyond H2S and NO interplay: hydrogen sulfide and nitroprusside react directly to give nitroxyl (HNO). A new pharmacological source of HNO. J Med Chem 56: 1499-1508, 2013. 24. Filipovic MR, Miljkovic J, Allgauer A, Chaurio R, Shubina T, Herrmann M, and IvanovicBurmazovic I. Biochemical insight into physiological effects of H(2)S: reaction with peroxynitrite and formation of a new nitric oxide donor, sulfinyl nitrite. Biochem J 441: 609-621, 2012. 25. Filipovic MR, Miljkovic J, Nauser T, Royzen M, Klos K, Shubina T, Koppenol WH, Lippard SJ, and Ivanovic-Burmazovic I. Chemical characterization of the smallest S-nitrosothiol, HSNO; cellular cross-talk of H2S and S-nitrosothiols. Journal of the American Chemical Society 134: 12016-12027, 2012. 26. Fox-Robichaud A, Payne D, Hasan SU, Ostrovsky L, Fairhead T, Reinhardt P, and Kubes P. Inhaled NO as a viable antiadhesive therapy for ischemia/reperfusion injury of distal microvascular beds. J Clin Invest 101: 2497-2505, 1998. 27. George TJ, Arnaoutakis GJ, Beaty CA, Jandu SK, Santhanam L, Berkowitz DE, and Shah AS. Inhaled hydrogen sulfide improves graft function in an experimental model of lung transplantation. The Journal of surgical research 178: 593-600, 2012. 28. Gherasim C, Yadav PK, Kabil O, Niu WN, and Banerjee R. Nitrite reductase activity and inhibition of H(2)S biogenesis by human cystathionine ss-synthase. PloS one 9: e85544, 2014. 29. Giustarini D, Rossi R, Milzani A, and Dalle‐Donne I. Nitrite and Nitrate Measurement by Griess Reagent in Human Plasma: Evaluation of Interferences and Standardization. Methods in enzymology 440: 361-380, 2008. 30. Gladwin MT, Schechter AN, Kim-Shapiro DB, Patel RP, Hogg N, Shiva S, Cannon RO, 3rd, Kelm M, Wink DA, Espey MG, Oldfield EH, Pluta RM, Freeman BA, Lancaster JR, Jr., Feelisch M, and Lundberg JO. The emerging biology of the nitrite anion. Nature chemical biology 1: 308-314, 2005. 31. Gladwin MT, Wang X, Reiter CD, Yang BK, Vivas EX, Bonaventura C, and Schechter AN. SNitrosohemoglobin is unstable in the reductive erythrocyte environment and lacks O2/NO-linked allosteric function. J Biol Chem 277: 27818-27828, 2002. 32. Gomes A, Fernandes E, and Lima JL. Use of fluorescence probes for detection of reactive nitrogen species: a review. J Fluoresc 16: 119-139, 2006. 33. Gow A, Doctor A, Mannick J, and Gaston B. S-Nitrosothiol measurements in biological systems. J Chromatogr B Analyt Technol Biomed Life Sci 851: 140-151, 2007.

19

662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708

34. Grisham MB, Johnson GG, and Lancaster JR, Jr. Quantitation of nitrate and nitrite in extracellular fluids. Methods in enzymology 268: 237-246, 1996. 35. Haouzi P, Chenuel B, Sonobe T, and Klingerman CM. Are H2S-trapping compounds pertinent to the treatment of sulfide poisoning? Clinical toxicology 52: 566, 2014. 36. Havenga MJ, van Dam B, Groot BS, Grimbergen JM, Valerio D, Bout A, and Quax PH. Simultaneous detection of NOS-3 protein expression and nitric oxide production using a flow cytometer. Analytical biochemistry 290: 283-291, 2001. 37. Hawkins CL and Davies MJ. Detection and characterisation of radicals in biological materials using EPR methodology. Biochimica et biophysica acta 1840: 708-721, 2014. 38. Heinrich TA, da Silva RS, Miranda KM, Switzer CH, Wink DA, and Fukuto JM. Biological nitric oxide signalling: chemistry and terminology. Br J Pharmacol 169: 1417-1429, 2013. 39. Hetrick EM and Schoenfisch MH. Analytical chemistry of nitric oxide. Annu Rev Anal Chem (Palo Alto Calif) 2: 409-433, 2009. 40. Honavar J, Bradley E, Bradley K, Oh JY, Vallejo MO, Kelley EE, Cantu-Medellin N, Doran S, Dell'italia LJ, Matalon S, and Patel RP. Chlorine gas exposure disrupts nitric oxide homeostasis in the pulmonary vasculature. Toxicology 321: 96-102, 2014. 41. Huajian T, Bo W, Kexin Z, Guangdong Y, Lingyun W, and Rui W. Oxygen-sensitive mitochondrial accumulation of cystathionine β-synthase mediated by Lon protease. Proceedings of the National Academy of Sciences of the United States of America, 2013. 42. Hunter CJ, Dejam A, Blood AB, Shields H, Kim-Shapiro DB, Machado RF, Tarekegn S, Mulla N, Hopper AO, Schechter AN, Power GG, and Gladwin MT. Inhaled nebulized nitrite is a hypoxia-sensitive NO-dependent selective pulmonary vasodilator. Nature medicine 10: 1122-1127, 2004. 43. Ida T, Sawa T, Ihara H, Tsuchiya Y, Watanabe Y, Kumagai Y, Suematsu M, Motohashi H, Fujii S, Matsunaga T, Yamamoto M, Ono K, Devarie-Baez NO, Xian M, Fukuto JM, and Akaike T. Reactive cysteine persulfides and S-polythiolation regulate oxidative stress and redox signaling. Proc Natl Acad Sci U S A 111: 7606-7611, 2014. 44. Ignarro LJ, Fukuto JM, Griscavage JM, Rogers NE, and Byrns RE. Oxidation of nitric oxide in aqueous solution to nitrite but not nitrate: comparison with enzymatically formed nitric oxide from Larginine. Proc Natl Acad Sci U S A 90: 8103-8107, 1993. 45. Jaffrey SR, Erdjument-Bromage H, Ferris CD, Tempst P, and Snyder SH. Protein S-nitrosylation: a physiological signal for neuronal nitric oxide. Nature cell biology 3: 193-197, 2001. 46. Janošík M, Kery V, Gaustadnes M, Maclean KN, and Kraus JP. Regulation of Human Cystathionine β-Synthase byS-Adenosyl-l-methionine: Evidence for Two Catalytically Active Conformations Involving an Autoinhibitory Domain in the C-Terminal Region†. Biochemistry 40: 1062510633, 2001. 47. Jhee KH and Kruger WD. The role of cystathionine beta-synthase in homocysteine metabolism. Antioxid Redox Signal 7: 813-822, 2005. 48. Jiang B, Tang G, Cao K, Wu L, and Wang R. Molecular mechanism for H(2)S-induced activation of K(ATP) channels. Antioxid Redox Signal 12: 1167-1178, 2010. 49. Kabil O and Banerjee R. Enzymology of H2S biogenesis, decay and signaling. Antioxid Redox Signal 20: 770-782, 2014. 50. Kabil O and Banerjee R. Redox biochemistry of hydrogen sulfide. J Biol Chem 285: 21903-21907, 2010. 51. Kabil O, Vitvitsky V, Xie P, and Banerjee R. The quantitative significance of the transsulfuration enzymes for H2S production in murine tissues. Antioxid Redox Signal 15: 363-372, 2011. 52. Kabil O, Zhou Y, and Banerjee R. Human cystathionine beta-synthase is a target for sumoylation. Biochemistry 45: 13528-13536, 2006.

20

709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 741 742 743 744 745 746 747 748 749 750 751 752 753 754 755 756

53. Kashfi K and Olson KR. Biology and therapeutic potential of hydrogen sulfide and hydrogen sulfide-releasing chimeras. Biochem Pharmacol 85: 689-703, 2013. 54. Kelleher ZT, Matsumoto A, Stamler JS, and Marshall HE. NOS2 regulation of NF-kappaB by Snitrosylation of p65. J Biol Chem 282: 30667-30672, 2007. 55. Kettenhofen NJ, Wang X, Gladwin MT, and Hogg N. In-gel detection of S-nitrosated proteins using fluorescence methods. Methods in enzymology 441: 53-71, 2008. 56. Kevil CG and Patel RP. S-Nitrosothiol biology and therapeutic potential in metabolic disease. Current opinion in investigational drugs 11: 1127-1134, 2010. 57. Kim-Shapiro DB and Gladwin MT. Pitfalls in measuring NO bioavailability using NOx. Nitric oxide : biology and chemistry / official journal of the Nitric Oxide Society 44C: 1-2, 2014. 58. Kim WS, Ye X, Rubakhin SS, and Sweedler JV. Measuring nitric oxide in single neurons by capillary electrophoresis with laser-induced fluorescence: use of ascorbate oxidase in diaminofluorescein measurements. Anal Chem 78: 1859-1865, 2006. 59. Klingerman CM, Trushin N, Prokopczyk B, and Haouzi P. H2S concentrations in the arterial blood during H2S administration in relation to its toxicity and effects on breathing. American journal of physiology Regulatory, integrative and comparative physiology 305: R630-638, 2013. 60. Koenitzer JR, Isbell TS, Patel HD, Benavides GA, Dickinson DA, Patel RP, Darley-Usmar VM, Lancaster JR, Jr., Doeller JE, and Kraus DW. Hydrogen sulfide mediates vasoactivity in an O2-dependent manner. American journal of physiology Heart and circulatory physiology 292: H1953-1960, 2007. 61. Kojima H, Nakatsubo N, Kikuchi K, Kawahara S, Kirino Y, Nagoshi H, Hirata Y, and Nagano T. Detection and imaging of nitric oxide with novel fluorescent indicators: diaminofluoresceins. Anal Chem 70: 2446-2453, 1998. 62. Kosower NS and Kosower EM. Thiol labeling with bromobimanes. Methods in enzymology 143: 76-84, 1987. 63. Krishnan N, Fu C, Pappin DJ, and Tonks NK. H2S-Induced sulfhydration of the phosphatase PTP1B and its role in the endoplasmic reticulum stress response. Sci Signal 4: ra86, 2011. 64. Lang JD, Jr., Smith AB, Brandon A, Bradley KM, Liu Y, Li W, Crowe DR, Jhala NC, Cross RC, Frenette L, Martay K, Vater YL, Vitin AA, Dembo GA, Dubay DA, Bynon JS, Szychowski JM, Reyes JD, Halldorson JB, Rayhill SC, Dick AA, Bakthavatsalam R, Brandenberger J, Broeckel-Elrod JA, Sissons-Ross L, Jordan T, Chen LY, Siriussawakul A, Eckhoff DE, and Patel RP. A randomized clinical trial testing the anti-inflammatory effects of preemptive inhaled nitric oxide in human liver transplantation. PLoS One 9: e86053, 2014. 65. Lang JD, Jr., Teng X, Chumley P, Crawford JH, Isbell TS, Chacko BK, Liu Y, Jhala N, Crowe DR, Smith AB, Cross RC, Frenette L, Kelley EE, Wilhite DW, Hall CR, Page GP, Fallon MB, Bynon JS, Eckhoff DE, and Patel RP. Inhaled NO accelerates restoration of liver function in adults following orthotopic liver transplantation. J Clin Invest 117: 2583-2591, 2007. 66. Lee Y, Oh BK, and Meyerhoff ME. Improved planar amperometric nitric oxide sensor based on platinized platinum anode. 1. Experimental results and theory when applied for monitoring NO release from diazeniumdiolate-doped polymeric films. Anal Chem 76: 536-544, 2004. 67. Lee ZW, Zhou J, Chen CS, Zhao Y, Tan CH, Li L, Moore PK, and Deng LW. The slow-releasing hydrogen sulfide donor, GYY4137, exhibits novel anti-cancer effects in vitro and in vivo. PloS one 6: e21077, 2011. 68. Li HD, Zhang ZR, Zhang QX, Qin ZC, He DM, and Chen JS. Treatment with exogenous hydrogen sulfide attenuates hyperoxia-induced acute lung injury in mice. European journal of applied physiology 113: 1555-1563, 2013. 69. Li L, Whiteman M, Guan YY, Neo KL, Cheng Y, Lee SW, Zhao Y, Baskar R, Tan CH, and Moore PK. Characterization of a novel, water-soluble hydrogen sulfide-releasing molecule (GYY4137): new insights into the biology of hydrogen sulfide. Circulation 117: 2351-2360, 2008.

21

757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802

70. Li Q and Lancaster JR, Jr. Chemical foundations of hydrogen sulfide biology. Nitric oxide : biology and chemistry / official journal of the Nitric Oxide Society 35: 21-34, 2013. 71. Liu J, Sun YQ, Zhang J, Yang T, Cao J, Zhang L, and Guo W. A ratiometric fluorescent probe for biological signaling molecule H2S: fast response and high selectivity. Chemistry 19: 4717-4722, 2013. 72. Lu C, Kavalier A, Lukyanov E, and Gross SS. S-sulfhydration/desulfhydration and Snitrosylation/denitrosylation: A common paradigm for gasotransmitter signaling by H< sub> 2 S and NO. Methods, 2013. 73. Luther GW, 3rd, Findlay AJ, Macdonald DJ, Owings SM, Hanson TE, Beinart RA, and Girguis PR. Thermodynamics and kinetics of sulfide oxidation by oxygen: a look at inorganically controlled reactions and biologically mediated processes in the environment. Front Microbiol 2: 62, 2011. 74. Madurga A, Mizikova I, Ruiz-Camp J, Vadasz I, Herold S, Mayer K, Fehrenbach H, Seeger W, and Morty RE. Systemic hydrogen sulfide administration partially restores normal alveolarization in an experimental animal model of bronchopulmonary dysplasia. American journal of physiology Lung cellular and molecular physiology 306: L684-697, 2014. 75. Mathai JC, Missner A, Kugler P, Saparov SM, Zeidel ML, Lee JK, and Pohl P. No facilitator required for membrane transport of hydrogen sulfide. Proc Natl Acad Sci U S A 106: 16633-16638, 2009. 76. Michael DP, Bailey TS, Matthew DH, and Leticia AM. Chemical Tools for Studying Biological Hydrogen Sulfide. In: Biochalcogen Chemistry: The Biological Chemistry of Sulfur, Selenium, and Tellurium: American Chemical Society, 2013, p. 15-32. 77. Minor RL, Jr., Myers PR, Guerra R, Jr., Bates JN, and Harrison DG. Diet-induced atherosclerosis increases the release of nitrogen oxides from rabbit aorta. J Clin Invest 86: 2109-2116, 1990. 78. Miyamoto R, Otsuguro K, Yamaguchi S, and Ito S. Contribution of cysteine aminotransferase and mercaptopyruvate sulfurtransferase to hydrogen sulfide production in peripheral neurons. J Neurochem 130: 29-40, 2014. 79. Montoya LA and Pluth MD. Selective turn-on fluorescent probes for imaging hydrogen sulfide in living cells. Chem Commun (Camb) 48: 4767-4769, 2012. 80. Munro AP and Williams DLH. Reactivity of sulfur nucleophiles towards S-nitrosothiols. Journal of the Chemical Society, Perkin Transactions 2: 1794-1797, 2000. 81. Mustafa AK, Gadalla MM, Sen N, Kim S, Mu W, Gazi SK, Barrow RK, Yang G, Wang R, and Snyder SH. H2S signals through protein S-sulfhydration. Sci Signal 2: ra72, 2009. 82. Mustafa AK, Sikka G, Gazi SK, Steppan J, Jung SM, Bhunia AK, Barodka VM, Gazi FK, Barrow RK, Wang R, Amzel LM, Berkowitz DE, and Snyder SH. Hydrogen sulfide as endothelium-derived hyperpolarizing factor sulfhydrates potassium channels. Circ Res 109: 1259-1268, 2011. 83. Nagahara N, Ito T, Kitamura H, and Nishino T. Tissue and subcellular distribution of mercaptopyruvate sulfurtransferase in the rat: confocal laser fluorescence and immunoelectron microscopic studies combined with biochemical analysis. Histochem Cell Biol 110: 243-250, 1998. 84. Ng ES, Jourd'heuil D, McCord JM, Hernandez D, Yasui M, Knight D, and Kubes P. Enhanced Snitroso-albumin formation from inhaled NO during ischemia/reperfusion. Circ Res 94: 559-565, 2004. 85. Nishida M, Sawa T, Kitajima N, Ono K, Inoue H, Ihara H, Motohashi H, Yamamoto M, Suematsu M, Kurose H, van der Vliet A, Freeman BA, Shibata T, Uchida K, Kumagai Y, and Akaike T. Hydrogen sulfide anion regulates redox signaling via electrophile sulfhydration. Nature chemical biology 8: 714724, 2012. 86. Niu WN, Yadav PK, Adamec J, and Banerjee R. S-Glutathionylation Enhances Human Cystathionine beta-Synthase Activity Under Oxidative Stress Conditions. Antioxid Redox Signal, 2014. 87. Nriagu JO. Stability of vivianite and ion-pair formation in the system Fe3(PO4)2-H3PO4-H2O. Geochimica et Cosmochimica Acta 36: 459-470, 1972.

22

803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835 836 837 838 839 840 841 842 843 844 845 846 847 848 849

88. Okamoto T, Tang X, Janocha A, Farver CF, Gladwin MT, and McCurry KR. Nebulized nitrite protects rat lung grafts from ischemia reperfusion injury. The Journal of thoracic and cardiovascular surgery 145: 1108-1116, 2013. 89. Olson KR. A practical look at the chemistry and biology of hydrogen sulfide. Antioxid Redox Signal 17: 32-44, 2012. 90. Olson KR, Deleon ER, Gao Y, Hurley K, Sadauskas V, Batz C, and Stoy GF. Thiosulfate: a readily accessible source of hydrogen sulfide in oxygen sensing. American journal of physiology Regulatory, integrative and comparative physiology 305: R592-603, 2013. 91. Ondrias K, Stasko A, Cacanyiova S, Sulova Z, Krizanova O, Kristek F, Malekova L, Knezl V, and Breier A. H(2)S and HS(-) donor NaHS releases nitric oxide from nitrosothiols, metal nitrosyl complex, brain homogenate and murine L1210 leukaemia cells. Pflugers Archiv : European journal of physiology 457: 271-279, 2008. 92. Pan J and Carroll KS. Persulfide reactivity in the detection of protein s-sulfhydration. ACS Chem Biol 8: 1110-1116, 2013. 93. Paul BD and Snyder SH. H(2)S signalling through protein sulfhydration and beyond. Nat Rev Mol Cell Biol 13: 499-507, 2012. 94. Perry MM, Hui CK, Whiteman M, Wood ME, Adcock I, Kirkham P, Michaeloudes C, and Chung KF. Hydrogen sulfide inhibits proliferation and release of IL-8 from human airway smooth muscle cells. Am J Respir Cell Mol Biol 45: 746-752, 2011. 95. Peter EA, Shen X, Shah SH, Pardue S, Glawe JD, Zhang WW, Reddy P, Akkus NI, Varma J, and Kevil CG. Plasma free H2S levels are elevated in patients with cardiovascular disease. J Am Heart Assoc 2: e000387, 2013. 96. Pickerodt PA, Emery MJ, Zarndt R, Martin W, Francis RC, Boemke W, and Swenson ER. Sodium nitrite mitigates ventilator-induced lung injury in rats. Anesthesiology 117: 592-601, 2012. 97. Poon BY, Raharjo E, Patel KD, Tavener S, and Kubes P. Complexity of inducible nitric oxide synthase: cellular source determines benefit versus toxicity. Circulation 108: 1107-1112, 2003. 98. Razavi HM, Wang L, Weicker S, Quinlan GJ, Mumby S, McCormack DG, and Mehta S. Pulmonary oxidant stress in murine sepsis is due to inflammatory cell nitric oxide. Crit Care Med 33: 1333-1339, 2005. 99. Razavi HM, Wang le F, Weicker S, Rohan M, Law C, McCormack DG, and Mehta S. Pulmonary neutrophil infiltration in murine sepsis: role of inducible nitric oxide synthase. Am J Respir Crit Care Med 170: 227-233, 2004. 100. Saito J, Mackay AJ, Rossios C, Gibeon D, Macedo P, Sinharay R, Bhavsar PK, Wedzicha JA, and Chung KF. Sputum-to-serum hydrogen sulfide ratio in COPD. Thorax 69: 903-909, 2014. 101. Samal AA, Honavar J, Brandon A, Bradley KM, Doran S, Liu Y, Dunaway C, Steele C, Postlethwait EM, Squadrito GL, Fanucchi MV, Matalon S, and Patel RP. Administration of nitrite after chlorine gas exposure prevents lung injury: effect of administration modality. Free radical biology & medicine 53: 1431-1439, 2012. 102. Searcy DG and Lee SH. Sulfur reduction by human erythrocytes. The Journal of Experimental Zoology 282: 310-322, 1998. 103. Sen N, Paul BD, Gadalla MM, Mustafa AK, Sen T, Xu R, Kim S, and Snyder SH. Hydrogen sulfidelinked sulfhydration of NF-kappaB mediates its antiapoptotic actions. Mol Cell 45: 13-24, 2012. 104. Shen X, Carlstrom M, Borniquel S, Jadert C, Kevil CG, and Lundberg JO. Microbial regulation of host hydrogen sulfide bioavailability and metabolism. Free radical biology & medicine 60: 195-200, 2013. 105. Shen X, Chakraborty S, Dugas TR, and Kevil CG. Hydrogen sulfide measurement using sulfide dibimane: critical evaluation with electrospray ion trap mass spectrometry. Nitric oxide : biology and chemistry / official journal of the Nitric Oxide Society 41: 97-104, 2014.

23

850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 869 870 871 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895

106. Shen X, Pattillo CB, Pardue S, Bir SC, Wang R, and Kevil CG. Measurement of plasma hydrogen sulfide in vivo and in vitro. Free radical biology & medicine 50: 1021-1031, 2011. 107. Shen X, Peter EA, Bir S, Wang R, and Kevil CG. Analytical measurement of discrete hydrogen sulfide pools in biological specimens. Free radical biology & medicine 52: 2276-2283, 2012. 108. Singh RJ, Hogg N, Joseph J, and Kalyanaraman B. Mechanism of nitric oxide release from Snitrosothiols. J Biol Chem 271: 18596-18603, 1996. 109. Spassov S, Pfeifer D, Strosing K, Ryter S, Hummel M, Faller S, and Hoetzel A. Genetic targets of hydrogen sulfide in ventilator-induced lung injury--a microarray study. PLoS One 9: e102401, 2014. 110. Stein A, Mao Z, Morrison JP, Fanucchi MV, Postlethwait EM, Patel RP, Kraus DW, Doeller JE, and Bailey SM. Metabolic and cardiac signaling effects of inhaled hydrogen sulfide and low oxygen in male rats. Journal of applied physiology 112: 1659-1669, 2012. 111. Sugimoto R, Okamoto T, Nakao A, Zhan J, Wang Y, Kohmoto J, Tokita D, Farver CF, Tarpey MM, Billiar TR, Gladwin MT, and McCurry KR. Nitrite reduces acute lung injury and improves survival in a rat lung transplantation model. American journal of transplantation : official journal of the American Society of Transplantation and the American Society of Transplant Surgeons 12: 2938-2948, 2012. 112. Szabo C, Coletta C, Chao C, Modis K, Szczesny B, Papapetropoulos A, and Hellmich MR. Tumorderived hydrogen sulfide, produced by cystathionine-beta-synthase, stimulates bioenergetics, cell proliferation, and angiogenesis in colon cancer. Proc Natl Acad Sci U S A 110: 12474-12479, 2013. 113. Szczesny B, Modis K, Yanagi K, Coletta C, Le Trionnaire S, Perry A, Wood ME, Whiteman M, and Szabo C. AP39, a novel mitochondria-targeted hydrogen sulfide donor, stimulates cellular bioenergetics, exerts cytoprotective effects and protects against the loss of mitochondrial DNA integrity in oxidatively stressed endothelial cells in vitro. Nitric oxide : biology and chemistry / official journal of the Nitric Oxide Society 41: 120-130, 2014. 114. Tao BB, Liu SY, Zhang CC, Fu W, Cai WJ, Wang Y, Shen Q, Wang MJ, Chen Y, Zhang LJ, Zhu YZ, and Zhu YC. VEGFR2 functions as an H2S-targeting receptor protein kinase with its novel Cys1045Cys1024 disulfide bond serving as a specific molecular switch for hydrogen sulfide actions in vascular endothelial cells. Antioxid Redox Signal 19: 448-464, 2013. 115. Taoka S and Banerjee R. Characterization of NO binding to human cystathionine beta-synthase: possible implications of the effects of CO and NO binding to the human enzyme. Journal of inorganic biochemistry, 2001. 116. Teng X, Scott Isbell T, Crawford JH, Bosworth CA, Giles GI, Koenitzer JR, Lancaster JR, Doeller JE, D WK, and R PP. Novel method for measuring S-nitrosothiols using hydrogen sulfide. Methods in enzymology 441: 161-172, 2008. 117. Terpolilli NA, Kim SW, Thal SC, Kataoka H, Zeisig V, Nitzsche B, Klaesner B, Zhu C, Schwarzmaier S, Meissner L, Mamrak U, Engel DC, Drzezga A, Patel RP, Blomgren K, Barthel H, Boltze J, Kuebler WM, and Plesnila N. Inhalation of nitric oxide prevents ischemic brain damage in experimental stroke by selective dilatation of collateral arterioles. Circ Res 110: 727-738, 2012. 118. Toubin C, Yeung DYH, English AM, and Peslherbe GH. Theoretical Evidence that Cu I Complexation Promotes Degradation of S-Nitrosothiols. Journal of the American Chemical Society 124: 14816-14817, 2002. 119. Tsukahara H, Ishida T, and Mayumi M. Gas-phase oxidation of nitric oxide: chemical kinetics and rate constant. Nitric oxide : biology and chemistry / official journal of the Nitric Oxide Society 3: 191198, 1999. 120. Vadivel A, Alphonse RS, Ionescu L, Machado DS, O'Reilly M, Eaton F, Haromy A, Michelakis ED, and Thebaud B. Exogenous hydrogen sulfide (H2S) protects alveolar growth in experimental O2-induced neonatal lung injury. PLoS One 9: e90965, 2014.

24

896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921 922 923 924 925 926 927 928 929 930 931 932 933 934 935 936 937 938 939 940 941 942

121. Verdon CP, Burton BA, and Prior RL. Sample pretreatment with nitrate reductase and glucose-6phosphate dehydrogenase quantitatively reduces nitrate while avoiding interference by NADP+ when the Griess reaction is used to assay for nitrite. Analytical biochemistry 224: 502-508, 1995. 122. Wang L, Frizzell SA, Zhao X, and Gladwin MT. Normoxic cyclic GMP-independent oxidative signaling by nitrite enhances airway epithelial cell proliferation and wound healing. Nitric oxide : biology and chemistry / official journal of the Nitric Oxide Society 26: 203-210, 2012. 123. Wang le F, Patel M, Razavi HM, Weicker S, Joseph MG, McCormack DG, and Mehta S. Role of inducible nitric oxide synthase in pulmonary microvascular protein leak in murine sepsis. Am J Respir Crit Care Med 165: 1634-1639, 2002. 124. Wang P, Zhang G, Wondimu T, Ross B, and Wang R. Hydrogen sulfide and asthma. Exp Physiol 96: 847-852, 2011. 125. Wang R. Physiological implications of hydrogen sulfide: a whiff exploration that blossomed. Physiol Rev 92: 791-896, 2012. 126. Wang R. Two's company, three's a crowd: can H2S be the third endogenous gaseous transmitter? FASEB journal : official publication of the Federation of American Societies for Experimental Biology 16: 1792-1798, 2002. 127. Wang X, Bryan NS, MacArthur PH, Rodriguez J, Gladwin MT, and Feelisch M. Measurement of nitric oxide levels in the red cell: validation of tri-iodide-based chemiluminescence with acidsulfanilamide pretreatment. J Biol Chem 281: 26994-27002, 2006. 128. Whiteman M, Li L, Kostetski I, Chu SH, Siau JL, Bhatia M, and Moore PK. Evidence for the formation of a novel nitrosothiol from the gaseous mediators nitric oxide and hydrogen sulphide. Biochemical and biophysical research communications 343: 303-310, 2006. 129. Whitfield NL, Kreimier EL, Verdial FC, Skovgaard N, and Olson KR. Reappraisal of H2S/sulfide concentration in vertebrate blood and its potential significance in ischemic preconditioning and vascular signaling. American journal of physiology Regulatory, integrative and comparative physiology 294: R1930-1937, 2008. 130. Wintner EA, Deckwerth TL, Langston W, Bengtsson A, Leviten D, Hill P, Insko MA, Dumpit R, VandenEkart E, Toombs CF, and Szabo C. A monobromobimane-based assay to measure the pharmacokinetic profile of reactive sulphide species in blood. Br J Pharmacol 160: 941-957, 2010. 131. Ya-Hong C, Rui W, Bin G, Yong-Fen Q, Pei-Pei W, Wan-Zhen Y, and Chao-Shu T. Endogenous hydrogen sulfide reduces airway inflammation and remodeling in a rat model of asthma. Cytokine, 2009. 132. Yadav AK, Doran SF, Samal AA, Sharma R, Vedagiri K, Postlethwait EM, Squadrito GL, Fanucchi MV, Roberts LJ, 2nd, Patel RP, and Matalon S. Mitigation of chlorine gas lung injury in rats by postexposure administration of sodium nitrite. American journal of physiology Lung cellular and molecular physiology 300: L362-369, 2011. 133. Yadav PK, Yamada K, Chiku T, Koutmos M, and Banerjee R. Structure and kinetic analysis of H2S production by human mercaptopyruvate sulfurtransferase. J Biol Chem 288: 20002-20013, 2013. 134. Yong QC, Hu LF, Wang S, Huang D, and Bian JS. Hydrogen sulfide interacts with nitric oxide in the heart: possible involvement of nitroxyl. Cardiovascular research 88: 482-491, 2010. 135. Zhang D, Macinkovic I, Devarie-Baez NO, Pan J, Park CM, Carroll KS, Filipovic MR, and Xian M. Detection of protein S-sulfhydration by a tag-switch technique. Angewandte Chemie 53: 575-581, 2014. 136. Zhang J, Wang X, Chen Y, and Yao W. Correlation between levels of exhaled hydrogen sulfide and airway inflammatory phenotype in patients with chronic persistent asthma. Respirology 19: 11651169, 2014. 137. Zhang P, Li F, Wiegman CH, Zhang M, Hong Y, Gong J, Chang Y, Zhang JJ, Adcock I, Chung KF, and Zhou X. Inhibitory Effect of Hydrogen Sulfide on Ozone-induced Airway Inflammation, Oxidative Stress and Bronchial Hyperresponsiveness. Am J Respir Cell Mol Biol, 2014.

25

943 944 945 946 947 948

138. Zhu W, Lin A, and Banerjee R. Kinetic properties of polymorphic variants and pathogenic mutants in human cystathionine gamma-lyase. Biochemistry 47: 6226-6232, 2008. 139. Zuckerbraun BS, Shiva S, Ifedigbo E, Mathier MA, Mollen KP, Rao J, Bauer PM, Choi JJ, Curtis E, Choi AM, and Gladwin MT. Nitrite potently inhibits hypoxic and inflammatory pulmonary arterial hypertension and smooth muscle proliferation via xanthine oxidoreductase-dependent nitric oxide generation. Circulation 121: 98-109, 2010.

949 950

Figure Legends:

951 952 953 954 955 956 957 958 959 960 961

Figure 1 NO-formation pathways: NO is synthesized by one of three nitric oxide synthases (NOS) that use L-arginine, oxygen and NADPH (reducing equivalents) as substrates to produce NO and citrulline. Citrulline is recycled back to L-Arginine via arginosuccinate synthase (ASS) and Arginosuccinate lyase (ASL). Alterantively, nitrite can be reduced to NO in a hypoxia and pH dependent manner via heme- or metalloproteins that can couple lowering oxygen tesnions and pH with a 1 electron reduction to NO. Once formed, NO can facilitate specific biological responses by nitrosylated heme proteins, by undergoing oxidation to facilitate nitrosation reactions resulting in protein S-nitrosothiols (RSNO) or nitrosamine (XNO) or undergoing rapid reactions with other free radicals (e.g. superoxide) to form peroxynitrite (ONOO-), which can mediate oxidation and nitration reactions. Ultimately, NO can be oxidized to nitrate, potentially via nitrite. Nitrate is then excreted or reduced back to nitrite by commensal nitrate-reducing bacteria in the oral cavity.

962 963 964 965 966 967 968 969 970 971

Figure 2 Sulfide metabolism and transsulfuration pathways CBS and CSE govern the flow of sulfur through homocysteine, cystathionine and cysteine, which is known as the transsulfuration pathway. These two enzymes are also the major catalytic pathways for generating H2S from cysteine and homocysteine in the cytosol. The third pathway via CAT and 3-MST may be more important in the mitochondrion and requires the transfer of sulfur from a persulfide (RSSH). On the other hand, persulfide can be a source of H2S at the expense of reductants such as glutathione (GSH) and thioredoxin (Trx(SH)2). In the mitochondrion, H2S reduces sulfide quinone oxidoreductase (SQR), generating a persulfide which is further transferred to an acceptor (RSH). This persulfide is then oxidized by persulfide dioxygenase (ETHE1) to SO32-, which is eventually converted to S2O32- and SO42- by rodanase and sulfite oxidase respectively.

972 973 974 975 976 977 978 979

Figure 3 Potential mechanisms by which H2S regulates protein thiol post-translational modifications S-nitrosation is a post-translation modification. In the illustrated scheme, the formation of SNO on a reactive thiol converts the protein from a “closed” conformation to an “open” one. SNO may serve as a target for H2S, which restores the protein to the “closed” form, generating thionitrous acid (HSNO) and hydrogen disulfide (HSSH) plus nitroxyl (HNO) in succession. However, HSNO may also mediate transnitrosation or undergo homolysis to NO, H2S or nitrite (26). HSSH may further modify the active thiol to a persulfide resulting in a stabilized “open” or “closed” conformation, which may vary from case to case.

980

26

981 982 983 984

27

985 986 987

28

988

29

Working with nitric oxide and hydrogen sulfide in biological systems.

Nitric oxide (NO) and hydrogen sulfide (H2S) are gasotransmitter molecules important in numerous physiological and pathological processes. Although th...
951KB Sizes 4 Downloads 11 Views